FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen,...

185
FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE RECYCLING AND MAGMA FORMATION BENEATH THE SOUTHERN CASCADE ARC by KRISTINA JANINE WALOWSKI A DISSERTATION Presented to the Department of Geological Sciences and the Graduate School of the University of Oregon in partial fulfillment of the requirements for the degree of Doctor of Philosophy June 2015

Transcript of FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen,...

Page 1: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE RECYCLING AND

MAGMA FORMATION BENEATH THE SOUTHERN CASCADE ARC

by

KRISTINA JANINE WALOWSKI

A DISSERTATION

Presented to the Department of Geological Sciences

and the Graduate School of the University of Oregon

in partial fulfillment of the requirements

for the degree of

Doctor of Philosophy

June 2015

Page 2: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

ii

DISSERTATION APPROVAL PAGE

Student: Kristina Janine Walowski

Title: From Cinder Cones to Subduction Zones: Volatile Recycling and Magma

Formation beneath the Southern Cascade Arc

This dissertation has been accepted and approved in partial fulfillment of the

requirements for the Doctor of Philosophy degree in the Department of Geological

Sciences by:

Dr. Paul Wallace Chairperson

Dr. Ilya Bindeman Core Member

Dr. Gene Humphreys Core Member

Dr. Darren Johnson Institutional Representative

and

Scott L. Pratt Dean of the Graduate School

Original approval signatures are on file with the University of Oregon Graduate School.

Degree awarded June 2015

Page 3: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

iii

© 2015 Kristina Janine Walowski

Page 4: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

iv

DISSERTATION ABSTRACT

Kristina Janine Walowski

Doctor of Philosophy

Department of Geological Sciences

June 2015

Title: From Cinder Cones to Subduction Zones: Volatile Recycling and Magma

Formation beneath the Southern Cascade Arc

Volatiles (H2O, CO2, S, Cl) play a key role in magmatic processes at subduction

zones. In this study, the dissolved volatile contents of olivine-hosted melt inclusions from

cinder cones in the Lassen segment of the Cascade arc are used to investigate dehydration

of subducted oceanic lithosphere, magma formation in the sub-arc mantle wedge, and

mafic magma storage and evolution in the crust.

Relatively young, hot oceanic lithosphere subducts beneath the Cascade arc. The

hydrogen-isotope and trace-element compositions of melt inclusions, when integrated

with thermo-petrologic modeling, demonstrate that fluids in Cascade magmas are sourced

from hydrated peridotite in the deep slab interior and that the oceanic crustal part of the

slab extensively dehydrates beneath the forearc. In contrast to their slab-derived H, the

melt inclusions have B concentrations and isotope ratios that are similar to mid-ocean

ridge basalt (MORB), requiring little to no slab contribution of B, which is also consistent

with extensive dehydration of the downgoing plate before it reaches sub-arc depths.

Correlations of volatile and trace element ratios in the melt inclusions (H2O/Ce, Cl/Nb,

Sr/Nd) demonstrate that geochemical variability in the magmas is the result of variable

amounts of addition of a hydrous subduction component to the mantle wedge. Radiogenic

Page 5: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

v

isotope ratios require that the subduction component has less radiogenic Sr and Pb and

more radiogenic Nd than the Lassen sub-arc mantle and is therefore likely to be a partial

melt of subducted Gorda MORB. These results provide evidence that chlorite-derived

fluids from the deep slab interior flux-melt the oceanic crust, producing hydrous slab

melts that migrate into the overlying mantle, where they react with peridotite to induce

further melting.

The basaltic magmas that erupted at Cinder Cone near Mt. Lassen trapped melt

inclusions during olivine crystallization at ~7-15 km depth. The melt inclusion

compositions require that two different mantle-derived magmas were involved in the

eruption, and temporal changes show that arrival of the two batches correlates with two

explosive phases of activity. Both magmas experienced rapid crustal contamination

before erupting, illustrating the complexities of cinder cone eruptions.

This dissertation includes previously published and unpublished co-authored

material.

Page 6: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

vi

CURRICULUM VITAE

NAME OF AUTHOR: Kristina Janine Walowski

GRADUATE AND UNDERGRADUATE SCHOOLS ATTENDED:

University of Oregon, Eugene

University of California, Los Angeles

DEGREES AWARDED:

Doctor of Philosophy, Geological Sciences, 2015, University of Oregon

Bachelor of Science, Geology, 2015, University of California, Los Angeles

AREAS OF SPECIAL INTEREST:

Geochemistry, Petrology, Volcanology

PROFESSIONAL EXPERIENCE:

Graduate Teaching Fellow (2010-2015).

University of Oregon, Department of Geological Sciences

Research Assistant (2011-2015)

University of Oregon, Department of Geological Sciences

Advisor: Paul Wallace

Undergraduate Research Assistant (2008-2010)

University of California, Los Angeles, Department of Earth and Space

Sciences

Advisor: Jeremy Boyce

GRANTS, AWARDS, AND HONORS:

GeoPRISMS AGU Student Prize for Best Oral Presentation (2014).

McMurray Award (2014). Department of Geological Science, University of

Oregon.

Goldschmidt Student Travel Grant (2013). Geochemical Society.

Graduate Research Award (2013). Graduate School, University of Oregon.

GeoPRISMS AGU Student Prize Honorable Mention (2012).

Staples Scholarship (2012). Department of Geological Science, University of

Oregon.

Valedictorian (2010), Department of Earth and Space Sciences, University of

California, Los Angeles.

Page 7: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

vii

PUBLICATIONS:

Walowski, K. J., Wallace, P. J., Hauri, E. H., Wada, I., & Clynne, M. A. (2015).

Slab melting beneath the Cascade Arc driven by dehydration of altered oceanic

peridotite. Nature Geoscience, 8(5), 404-408.

Walowski K, Wallace P, Clynne M. (2014). Slab Melting and Magma Formation

beneath the southern Cascade Arc. AGU Fall Meeting Abstracts, V31G-03.

Walowski K, Wallace P, Hauri E, Clynne M & Wada I (2014). Slab Dehydration

beneath the Southern Cascade Arc Inferred from B and H Isotopes. Goldschmidt Abstracts,

2014, 2613.

Walowski, K. J., Wallace, P. J., Hauri, E. H., Clynne, M. A., Rea, J., &

Rasmussen, D. J. (2013). Magma formation in hot-slab subduction zones: Insights from

hydrogen isotopes in Cascade Arc melt inclusions. In AGU Fall Meeting Abstracts (Vol.

1, p. 01).

Walowski K, Wallace P, Clynne M, Wada I & Rasmussen D (2013). Magma

Formation in Hot-Slab Subduction Zones: Insights from Volatile Contents of Melt Inclusions

from the Southern Cascade Arc. Mineralogical Magazine, 77(5), 2441.

Walowski K, Wallace P, Clynne M, Wada I & Rasmussen D (2012). Understanding magma formation and mantle conditions in the Lassen segment of the Cascade

Arc: Insights from volatile contents of olivine-hosted melt inclusions. AGU Fall Meeting

Abstracts, V14B-01.

Walowski, K. J., Wallace, P. J., Cashman, K. V., & Clynne, M. A. (2011).

Inferring the magmatic plumbing system and melt evolution from olivine-hosted melt

inclusions at Cinder Cone, Lassen Volcanic National Park, California. In AGU Fall

Meeting Abstracts (Vol. 1, p. 02).

Walowski, K. J., and Boyce, J.W. (2009). Volatile History of the Bishop Tuff from

Apatite Phenocrysts and Inclusions. AGU Fall Meeting Abstracts,V23C-2093.

Page 8: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

viii

ACKNOWLEDGMENTS

This dissertation has benefited from the contributions of many people. First, I

would like to thank my advisor, Paul Wallace, for his guidance and support throughout

my Ph.D. I can’t imagine a better Ph.D. advisor, and I am privileged to have gotten the

opportunity to work with him. I’d also like to thank the faculty at the University of

Oregon, and specifically, my dissertation committee members, Ilya Bindeman for

teaching me about isotopes, Gene Humphreys for helping me see the bigger picture, and

Darren Johnson. My dissertation benefited from a number of collaborations, and includes

the contributions of many co-authors, apart from Paul and myself. Thanks to Erik Hauri,

Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and

Philipp Ruprecht for their individual contributions, and John Donavan, Adam Kent, and

Brian Monteleone for their assistance with various analytical instruments. In particular, I

would like to thank Mike Clynne, whose Ph.D. thesis and continued work in the Lassen

area laid the foundation my dissertation research, and for his important contributions to

all three chapters.

This work was largely funded by the National Science Foundation through research

grants awards to Paul Wallace and Kathy Cashman (EAR-1019848 and EAR-1119224).

I’d also like to thank the incredible graduate community that I was so lucky be a

part of. To the few who were there from start to finish – Kristin Sweeney, Corina

Cerovski-Darriau, and Scott Maguffin – we’ve had a great run, and I hope it doesn’t end

here! Special thanks to Angie Seligman, who the first floor would not have been the same

without, Randy Krogstad, who will always be one of my best friends, Amberlee Darold

and Gillean Arnoux, who made Cascade 124 a better place, and, of course, Captain

Page 9: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

ix

Malcom Reynolds.

I’d also like to specially thank my first academic ‘dad’, Jeremy Boyce. He inspired

me to follow this path and prepared me so well for the journey that has been my Ph.D.

Five years later, he is still there for me whenever I need advice and guidance, which is

something I truly value.

To my Mom and Dad, I could never thank you enough for your unconditional love

and support. You have both taught me so much, and given me everything I could have

ever needed. It is difficult to put into words how much you both mean to me, and I love

you both very much. To my brother, Chris, thank you for showing me the coolness of

science. I truly value our friendship and look forward to joining you in the higher-ed

scientist club. Finally, I’d like to thank my partner Alexander Handwerger, without

whom none of this would have been possible. Thank you for being there for me every

single step of the way, and bringing me so much happiness. Our time in Oregon has been

incredible, and I can’t wait to start our next adventure together in Scotland!

Page 10: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

x

To My Family

Page 11: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xi

TABLE OF CONTENTS

Chapter Page

I. INTRODUCTION .................................................................................................... 1

1.1. Water, Subduction, and the Cascade Arc ........................................................ 1

1.2. Scientific Approach ........................................................................................ 2

1.3. Dissertation Organization ............................................................................... 3

II. SLAB MELTING BENEATH THE CASCADE ARC DRIVEN BY

DEHYDRATION OF ALTERED OCEANIC PERIDOTITE .............................. 7

2.1. Introduction ..................................................................................................... 7

2.2. Slab Surface Temperatures ............................................................................. 11

2.3. Hydrogen Isotopes .......................................................................................... 13

2.4. D/H Fractionation Model ................................................................................ 15

2.5. Evidence and Implications for Slab Melting .................................................. 18

2.6. Bridge .............................................................................................................. 20

III. SLAB MELTING AND MAGMA FORMATION BENEATH THE

SOUTHREN CASCADE ARC ............................................................................. 21

3.1. Introduction ..................................................................................................... 21

3.2. Geologic Setting.............................................................................................. 24

3.3. Sample Descriptions and Analytical Methods ................................................ 26

3.4. Results ............................................................................................................. 28

3.4.1. Melt Inclusion Major and Trace Element Compositions ....................... 28

3.4.2. Magmatic Volatile Contents .................................................................. 31

3.4.3. Isotopic Compositions ........................................................................... 32

3.5. Discussion ....................................................................................................... 33

Page 12: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xii

Chapter Page

3.5.1. The Source of Volatiles in the Lassen Region Primitive Magmas ........ 33

3.5.2. Evidence for Slab Melting .................................................................... 37

3.5.3. The Role of Sediment Melts and Crustal Assimilation ......................... 41

3.5.4. Modeling Slab Melt Addition to the Mantle Wedge ............................. 43

3.5.5. The Sr/Y Adakite Signature ................................................................... 47

3.5.6. Model for Melt Generation beneath the Southern Cascade Arc ............ 49

3.6. Conclusions ..................................................................................................... 52

3.7. Bridge .............................................................................................................. 53

IV. UNDERSTANDING MELT EVOLUTION AND THE MAGMATIC

PLUMBING SYSTEM AT CINDER CONE, LASSEN VOLCANIC

NATIONAL PARK CALIFORNIA: INSIGHTS FROM

OLIVINE-HOSTED MELT INCLUSIONS.......................................................... 54

4.1. Introduction ..................................................................................................... 54

4.2. Geologic Setting.............................................................................................. 58

4.3. Sample Descriptions and Tephra Stratigraphy ............................................... 59

4.4. Methods........................................................................................................... 63

4.4.1. Sample Preparation ................................................................................ 63

4.4.2. Analytical Procedures ............................................................................ 64

4.5. Results ............................................................................................................. 66

4.5.1. Olivine Compositions ............................................................................ 66

4.5.2. Post-entrapment Crystallization and Fe-loss Corrections ...................... 67

4.5.3. Major Element Compositions ................................................................ 68

4.5.4. Volatile Concentrations ......................................................................... 70

Page 13: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xiii

Chapter Page

4.5.5. Trace Element Concentrations ........................................................................... 75

4.6. Discussion ....................................................................................................... 76

4.6.1. Crustal Contamination at Cinder Cone .................................................. 76

4.6.2. Parental Melt Compositions ................................................................... 79

4.6.3. Timescales of Mixing and Ascent.......................................................... 84

4.6.4. Mechanisms of Crustal Contamination and

Plumbing System Evolution ................................................................ 87

4.7. Conclusions ..................................................................................................... 91

APPENDICES ............................................................................................................. 93

A. CHAPTER II SUPPLEMENTARY MATERIALS .......................................... 93

A.1. Methods .............................................................................................. 93

A.2. Supplementary Discussion ................................................................. 95

A.3. Hydrogen Isotope and Trace Element Compositions ........................ 113

B. CHAPTER III SUPPLEMENTARY MATERIALS......................................... 125

B.1. Restored Melt Inclusion Compositions .............................................. 125

B.2. Boron Isotope Compositions .............................................................. 134

B.3. Radiogenic Isotope Compositions ...................................................... 135

C. CHAPTER IV SUPPLEMENTARY MATERIALS ........................................ 136

C.1. PEC Corrected Melt Inclusion Compositions .................................... 136

C.2. Analyzed (uncorrected) Melt Inclusion Compositions ...................... 143

REFERENCES CITED ................................................................................................ 149

Page 14: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xiv

LIST OF FIGURES

Figure Page

2.1 Cascade Arc Regional Map and Lassen Sample Locations ................................... 9

2.2 Melt Inclusion Compositions and Slab Surface Temperatures .............................. 13

2.3 Hydrogen Isotopes: Measured Values and Model Results .................................... 17

3.1 Sample Locations and Across-Arc Geochemical Variability ................................ 25

3.2 Corrected Melt Inclusion Major and Trace Element Compositions ...................... 30

3.3 Boron Isotope Compositions.................................................................................. 33

3.4 Corrected Melt Inclusion Volatile and Trace Element Ratios ............................... 36

3.5 Radiogenic Isotope Compositions and Slab Melt Mixing Models ........................ 39

3.6 Percent Melt and source H2O: Results from pMELTS .......................................... 44

3.7 Comparison of pMELTS Model Results and Melt Inclusion Compositions ......... 46

3.8 The Effect of Slab Melt Addition to Peridotite for Sr/Y, Y and H2O .................... 48

3.9 Schematic Interpretation for Magma Formation beneath the S. Cascades ............ 52

4.1 Geologic and Isopach Map of Cinder Cone ........................................................... 60

4.2 Stratigraphic Section of Sample Locality .............................................................. 62

4.3 Olivine Phenocryst Compositions.......................................................................... 67

4.4 Melt Inclusion Total Alkalis and SiO2 Compared to Bulk Tephra and Lava ........ 70

4.5 Melt Inclusion MgO vs. Al2O3 and CaO ............................................................... 71

4.6 Melt Inclusion Corrected H2O vs. CO2 .................................................................. 75

4.7 Spider Diagram of Average Melt Inclusions Compositions .................................. 76

4.8 Modeled Mixing with Granitic Xenoliths .............................................................. 78

Page 15: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xv

Figure Page

4.9 Melt Inclusion Li vs. Pb ......................................................................................... 80

4.10 Melt Inclusion MgO vs. Trace Elements ............................................................. 82

4.11 Comparison of Cinder Cone to Lassen Region Magmas ..................................... 84

4.12 Ni Diffusion in Olivine ........................................................................................ 86

4.13 Temporal Variability in Bulk Tephra Composition ............................................. 89

4.14 Core to Rim Variations in Olivine Xenocrysts .................................................... 90

Page 16: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

xvi

LIST OF TABLES

Table Page

2.1 Average Volatile, Trace Element, and Stable Isotope Compositions of

Melt Inclusions from the Lassen Region Cinder Cones ........................................ 11

3.1 Averaged Primary Melt Compositions .................................................................. 29

4.1 Tephra Stratigraphy and Sample Descriptions ...................................................... 63

Page 17: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

1

CHAPTER I

INTRODUCTION

1.1. Water, Subduction, and the Cascade Arc

The distribution of water in the Earth’s mantle has a strong impact on the

evolution of our planet. Introduced to the mantle through subduction, water makes active

plate tectonics possible on Earth by reducing rock viscosities, facilitating mantle

convection (e.g. Jung and Karato, 2001). Melting of the mantle wedge in subduction

zones would not be possible without the addition of water (e.g., Ringwood, 1974), and

once these melts reach the surface, their water contents directly affect the explosivity of

the magma (e.g., Roggensack, 1997; Metrich and Wallace, 2008). The influence water

has on a number of processes, such as the formation of volcanic arcs and production of

continental crust, makes understanding the origins and distribution of water in subduction

zones a very important area of research in the earth sciences.

Perhaps the most important role of H2O in subduction zones is driving magma

production. When the oceanic lithosphere is subducted, increasing temperatures and

pressures cause the progressive breakdown of hydrous phases in a series of reactions that

release H2O into the overlying mantle wedge (e.g., Schmidt and Poli, 1998). In the classic

model of petrogenesis at subduction zones, these aqueous fluids infiltrate the mantle

wedge, lower the solidus, and cause melting. However, convergent margins vary globally

- they subduct oceanic lithosphere of different ages and at different rates (e.g., Syracuse

et al., 2010), and have sediment packages of different thicknesses and compositions (e.g.,

Plank & Langmuir, 1993). Different combinations of plate age and convergence rate

Page 18: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

2

control how rapidly temperatures in the slab increase with depth, and therefore influence

the depths at which dehydration reactions occur (e.g., Schmidt & Poli, 1998; Hacker,

2008), and the composition of the fluids and/or melts that infiltrate the mantle wedge

(e.g., Kessel et al., 2005).

The Cascadia subduction zone, which runs from northern California to southern

British Columbia, represents a global endmember, subducting very young oceanic

lithosphere (6-10 Ma at the trench; Wilson et al., 2002, Kirby et al., 2002). The processes

that drive magma formation beneath the Cascade arc and other warm-slab subduction

zones have been debated because young oceanic crust should largely dehydrate beneath

the forearc (Van Keken et al., 2010), where the mantle wedge is not at high enough

temperatures to melt. In addition, geochemical variability along strike in the Cascades has

led to contrasting interpretations about the role of volatiles in magma generation (e.g.,

Leeman et al., 1990; Grove et al., 2002; Ruscitto et al., 2010). In this dissertation, I focus

on the Lassen segment of the Cascade arc, where previous studies have demonstrated that

across-arc geochemical variations in primitive basaltic magmas result from varying

extents of mantle wedge fluxing by material recycled from the slab. Using direct

measurements of dissolved magmatic volatiles and a number of other geochemical tools,

I show how the subduction of warm oceanic lithosphere beneath the Cascades influences

dehydration reactions in the downgoing slab and melt production in the mantle wedge.

1.2. Scientific Approach

Volatiles, such as H2O, CO2, S, Cl, and F, play a key role a number magmatic

processes. Pre-eruptive volatile contents are difficult to estimate, however, because they

Page 19: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

3

decrease in solubility as pressure decreases and exsolve from the magma during ascent,

driving explosive eruptions (e.g., Roggensack, 1997; Metrich and Wallace, 2008). As a

result, most magmatic volatiles are lost to the atmosphere, leaving the resulting volcanic

material with only a small fraction of the volatile contents with which it began (e.g.,

Holloway and Blank, 1994). In this dissertation, I utilize the geochemistry of melt

inclusions hosted in olivine crystals to get around this problem. Melt inclusions are small

volumes of melt trapped inside of phenocrysts at depth. Olivine-hosted melt inclusions

are one of the best tools currently available to directly measure pre-eruptive volatile

concentrations, because they are often trapped at depth, before significant degassing

occurs (Anderson, 1974; Schiano, 2003). Because they are trapped at depth, they not

only provide estimates of pre-eruptive volatiles, but may also provide a unique snapshot

of melt compositions, before magmatic evolution occurs.

1.3. Dissertation Organization

This dissertation is separated into three stand-alone chapters that are prepared as

individual manuscripts for publication. Chapters II and III investigate deep magmatic

processes, such as mantle melting and slab recycling, while Chapter IV focuses on

shallow (crustal level) magmatic processes that occur prior to and during cinder cone

eruptions. The unifying theme of the dissertation is the use of melt inclusions in olivine

sampled from tephra deposits at mafic cinder cones. Throughout the dissertation, I show

that melt inclusion datasets can be utilized to answer a wide range of questions when

integrated with other types of data (e.g., radiogenic isotopes, mineral chemistry, eruptive

Page 20: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

4

stratigraphy) and modeling (e.g., geodynamic, geochemical, and thermodynamic

modeling).

In Chapter II, “Slab Melting beneath the Cascade Arc driven by dehydration of

altered oceanic peridotite,” I use the hydrogen isotope composition of basaltic melt

inclusions from seven different cinder cones to fingerprint the fluid sources that initiate

magma generation beneath the Cascade arc. I also develop a thermo-petrologic model to

calculate progressive D/H fractionation in the subducting plate to inform the geochemical

data. The results show that the downgoing plate beneath the southern Cascades

experiences extensive dehydration and that fluids derived from the slab interior may flux-

melt the slab top. In Chapter III, “Slab Melting and Magma Formation beneath the

Southern Cascade Arc,” I use a larger melt inclusion dataset to further test the

interpretations presented in Chapter II. I use B concentrations and isotope ratios, trace

elements, and the volatile concentrations of the melt inclusions along with the radiogenic

isotope ratios of bulk tephra from the eight different cinder cones (one additional cone to

the seven discussed in Chapter II). This dataset is used to demonstrate that geochemical

variability of primitive basalts in the Lassen region is the result of variable amounts of

addition of a hydrous subduction component to the mantle wedge, and also provides

strong evidence that melting of the subducted basaltic oceanic crust beneath the Lassen

region plays an important role in magma formation in the mantle wedge. The combined

results of Chapters II and III indicate that chlorite-derived fluids from the deep slab

interior can flux-melt the oceanic crust, producing hydrous slab melts that migrate into

the overlying mantle, where they react with peridotite to induce further melting. I suggest

that this mechanism for slab melting and magma formation may be common in arcs

Page 21: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

5

beneath which young oceanic crust subducts, such as central Mexico (e.g., Cai et al.,

2014).

In Chapter IV, I investigate the compositional evolution of magmas from the

youngest cinder cone in the Cascade arc, Cinder Cone (eruption age 1666 C.E.; Sheppard

et al., 2009). I integrate melt inclusion compositions with stratigraphic information, bulk

lava and tephra compositions, mineral chemistry, and physical volcanology to better

understand the processes and timescales associated with the storage, ascent, and eruption

of magmas at this monogenetic volcano. I find that two different mantle-derived magmas

were tapped during the eruption of Cinder Cone, correlating with two explosive eruptive

phases. Both magmas experienced rapid crustal contamination before erupting at the

surface similar to other well-documented cinder cone eruptions (e.g., Paricutin and

Jorullo, in Meixco; e.g., Pioli et al., 2008, Johnson et al., 2008). However, the timescales

and mechanisms of contamination and eruption are shown to occur more rapidly at

Cinder Cone than monogenetic volcanoes in central Mexico, illustrating the complexities

of short-lived cinder cone eruptions.

For all chapters in this dissertation, I am the first author, and performed all of the

laboratory work, field-based work, and analytical work, unless explicitly stated. Chapter

II has been published in Nature Geoscience with Dr. Paul Wallace (advisor), Dr. Erik

Hauri (Carnegie Instiution of Washington), Dr. Ikuko Wada (University of Minnesota),

and Dr. Michael Clynne (U.S. Geological Survey). Chapter III is in preparation in the

style of Earth and Planetary Science Letters and will be coauthored by Dr. Paul Wallace

(advisor), Dr. Michael Clynne (U.S. Geological Survey), Dan Rassmussen (Ph.D.

candidate, Lamont-Doherty Earth Observatory), and Dr. Dominique Weis (University of

Page 22: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

6

British Columbia). Chapter IV is in preparation in the style of Bulletin of Volcanology

and will be coauthored by Dr. Paul Wallace (advisor), Dr. Michael Clynne (U.S.

Geological Survey), Dr. Kathy Cashman (University of Bristol), J. Katie Marks (M.S.,

University of Oregon), and Dr. Philipp Ruprecht (Lamont-Doherty Earth Observatory).

Page 23: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

7

CHAPTER II

SLAB MELTING BENEATH THE CASCADE ARC DRIVEN BY

DEHYDRATION OF ALTERED OCEANIC PERIDOTITE

This work was published in volume 8 of the journal Nature Geoscience in May,

2015. Sample collection was performed by me with the help of Dr. Wallace and Dr.

Clynne. Sample preparation for and Nano-secondary ion mass spectrometry volatile and

hydrogen isotope analyses were performed by me with the help of Dr. Hauri. Data

corrections, analyses, and interpretation were performed by me, with the help of

Dr.Wallace. The fractionation model was created by me with the help of Dr. Wada and

incorporated her previously published work. I drafted all the figures used in this article

and the text was written by me, with editorial assistance by Dr. Wallace.

2.1. Introduction

Water is returned to Earth’s interior at subduction zones. However, the processes

and pathways by which water leaves the subducting plate and causes melting beneath

volcanic arcs are complex; the source of the water—subducting sediment, altered oceanic

crust, or hydrated mantle in the downgoing plate—is debated; and the role of slab

temperature is unclear. Here we analyse the hydrogen-isotope and trace-element

signature of melt inclusions in ash samples from the Cascade Arc, where young, hot

lithosphere subducts. Comparing these data with published analyses, we find that fluids

Page 24: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

8

in the Cascade magmas are sourced from deeper parts of the subducting slab—hydrated

mantle peridotite in the slab interior—compared with fluids in magmas from the

Marianas Arc, where older, colder lithosphere subducts. We use geodynamic modelling

to show that, in the hotter subduction zone, the upper crust of the subducting slab rapidly

dehydrates at shallow depths. With continued subduction, fluids released from the deeper

plate interior migrate into the dehydrated parts, causing those to melt. These melts in turn

migrate into the overlying mantle wedge, where they trigger further melting. Our results

provide a physical model to explain melting of the subducted plate and mass transfer

from the slab to the mantle beneath arcs where relatively young oceanic lithosphere is

subducted.

The Cascadia subduction zone and associated magmatic arc in the U.S. and

Canada is a global endmember example of a “warm-slab” subduction zone 1,2

, in which

very young (6-9 Ma at the trench)3 and therefore hot oceanic crust subducts (Fig. 2.1).

Such arcs have proven to be a challenge to our current understanding of arc magmatism

because the hotter nature of the young subducted plate causes most dehydration of the

oceanic crust to occur beneath the forearc rather than the arc4,5

. As a result, little H2O

may reach sub-arc depths to initiate melt production in the overlying mantle wedge,

unless there is substantial H2O carried to greater depths by hydrated (serpentinized)

peridotite in the mantle portion of the downgoing slab6. Basaltic magmas in the Cascades

have H2O contents (3.2 ±1.2 wt%)7,8,9

that are only slightly lower than the global arc

average (3.9±0.45 wt%, 1 s.d.)10

, raising questions about the extent of dehydration of the

subducting plate beneath the forearc and the extent to which hydrated mantle in the

subducting plate carries H2O to sub-arc depths.

Page 25: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

9

Figure 2.1. Cascade arc regional map and Lassen sample locations. a) Regional map of

the Northwestern United States showing major tectonic boundaries. Black arrows show

the direction and convergence rate of the Juan de Fuca and Gorda plates. Lassen Peak

(red triangle) represents the southernmost active volcanic center of the Quaternary

Cascade arc. b) Expanded map of the Lassen Region with locations of Quaternary cinder

cones sampled in this study. Symbols and colors for calc-alkaline basalt (CABs; squares)

and CAB-transitional (circles) samples are used throughout the article.

Hydrogen isotopes have a larger mass difference than any other isotope pair, and

as a result, they exhibit strong isotopic fractionation during various geologic processes. In

subduction zones, as hydrous minerals break down with increasing temperature and

Page 26: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

10

pressure, the D/H ratio is fractionated, producing D-enriched fluids and a complementary

residual slab that is isotopically lighter11

. Recent work in the Marianas has shown that

because the contribution of D and H from the overlying mantle is minimal, the δD

signature (δD=1,000×[(D/H)sample − (D/H)SMOW] /(D/H)SMOW, where SMOW is standard

mean ocean water with δD=0) of slab fluids may be retained in primitive arc magmas11

.

This makes magmatic D/H ratios a valuable tool for investigating volatile recycling in the

Cascade arc because the thermal structure of the slab strongly influences the depths at

which key dehydration reactions occur in the downgoing plate6,12

.

Here we focus on the Lassen segment of the Cascade arc, the southernmost active

segment of the arc, where previous work has shown across-arc geochemical variations

related to subduction enrichment13,14

(Fig. 2.1a). Our samples were collected from tephra

deposits associated with primitive basaltic cinder cones (MgO > 7 wt%; Supplementary

Table A.1; see Appendix A for all supplementary information for this chapter), which

span ~80 km across the arc axis (Fig. 2.1b). Melt inclusions – glassy “blobs” of silicate

melt trapped inside olivine crystals – were used in this study to measure pre-eruptive

volatile contents. We present the analyses of melt inclusions from four calc-alkaline

basalts (CAB) and three basalts that are transitional between CAB and low-K tholeiite

(LKT) based on trace element patterns (Supplementary Fig. A.1). Using the maximum

volatile contents at each cone to represent the undegassed magma, we find values of 2.47-

3.4 wt% H2O for CABs and 1.3-1.4 wt% H2O for transitional CABs in the Lassen region

(Table 2.1). The CAB values are similar to H2O contents of other Cascade CAB

magmas7,8,9

, whereas the transitional CABs have lower values, as expected, given that

Page 27: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

11

they are transitional towards LKTs, which are interpreted to be relatively dry magmas

formed by decompression melting15

.

Table 2.1. Average volatile, trace element and stable isotope composition of melt

inclusions from Lassen Region cinder cones.

In addition to H2O, trace element ratios of large ion lithophile elements (LILE)

and light rare earth elements (LREE) are also commonly used to distinguish addition of a

hydrous subduction component to the mantle source because of the fluid mobility of

LILEs and relative depletion of LREE in the mantle wedge. In the Lassen region, (Sr/P)N

(N refers to normalization to primitive mantle16

) has been used to distinguish a

subduction component because it correlates with LILE/LREE ratios and varies

systematically across the arc13,14

. For magmas from the Lassen region, (Sr/P)N correlates

with other slab component tracers such as H2O/Ce (Fig. 2.2a), demonstrating a clear link

between volatile and trace element enrichment of the mantle wedge as the result of

variable addition of a hydrous slab component.

2.2. Slab Surface Temperatures

To better understand thermal conditions at the slab-wedge interface, we

developed a 2-D steady-state thermal model for this region2. The maximum depth of slab-

Page 28: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

12

mantle decoupling (MDD) controls the trench-ward extent of solid mantle wedge flow,

and this depth tends to be 70-80 km for most subduction zones2. In our model for the

Lassen segment, we assume that the MDD is 75 km. Temporal changes in regional

tectonics or slab geometry can cause deviation from the common MDD. Thus, the range

of slab surface temperatures predicted by models with MDD of 65–85 km depth are

shown in Figure 2.2b. We also calculated slab surface temperatures using the H2O/Ce

ratio17

(see Supplementary Information A.) as an independent comparison with the

thermal model. Slab surface temperatures at sub-arc depths of ~85-95 km18

beneath the

Lassen segment calculated using this method are 725-850 ± 50 °C (Fig. 2.2b), slightly

lower than temperatures predicted by the thermal model. Our thermal model does not

include the effects of fluid circulation within the oceanic crust at shallow depths in the

subduction zone or the latent heat of fusion that would affect temperatures if the oceanic

crust at the plate top was partially melted (see Supplementary Information A.2). Each of

these effects could reduce the slab surface temperatures by ~50 °C19

. Furthermore, in

interpreting H2O/Ce temperatures, we neglect the potential effects of lateral migration of

fluids and melts through the wedge20

which may contribute to the difference between the

two estimates (see Supplementary Information A.2). Regardless of the small offset, both

the thermal model and H2O/Ce temperature estimates indicate active fluxing of hydrous

material from the slab into the mantle wedge beneath the arc rather than down-dragging

of hydrous mantle from the forearc region. If the latter was an important process, it

would have imprinted the mantle wedge with relatively high H2O/Ce, resulting in lower

apparent temperatures17

. Calculated slab surface temperatures beneath the Cascades

(from both methods) also plot at or above the solidi of both mid-ocean ridge basalt

Page 29: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

13

(MORB)+H2O21

and sediment+H2O22

(Fig. 2.2b), indicating the likelihood of partial

melting of the slab top if H2O is present.

Figure 2.2. Melt inclusion compositions and slab surface temperatures. a) H2Omax/Ce

versus (Sr/P)N for least degassed melt inclusions. 10% partial melts of depleted mantle

(DMM)46

and more enriched mantle47

(E-M; 66% DMM and 33% E-M; see

Supplementary Fig. A.2 and Information) mixed with a hydrous component were used to

calculate curves, labeled with the % of total H2O contributed by the subduction

component in the final magma. b) Slab surface temperatures from the H2O/Ce

thermometer17

, 2-D thermal model for the Lassen segment (MDD of 75 km; solid black

curve), and the sediment + H2O 22

and MORB + H2O solidi21

. Temperatures predicted by

thermal models with MDDs of 65–85 km shown by grey shaded region.

2.3. Hydrogen Isotopes

The melt inclusion data from the Lassen region show a negative correlation

between measured δD values and H2O (Supplementary Fig. A.3). Recent experimental

work demonstrates that this correlation is the result of post-entrapment fractionation

caused by preferential diffusive loss of H relative to D through the olivine host23, 24

, and

Page 30: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

14

our data show the first evidence for this process in naturally glassy samples

(Supplementary Fig. A.3). Using a diffusive re-equilibration model24

, we calculated the

initial δD composition for each cone by assuming that the highest H2O concentrations

measured are in melt inclusions that are least affected by this process (Supplementary

Fig. A.3). We compare our results (Fig. 2.3a) with published data from the Marianas11

where negative correlations between δD and H2O wt% are not observed. The Marianas

data come from longer-lived volcanoes that erupted more explosively, and diffusive H

loss may have been minimized by rapid ascent rates and lower magmatic temperatures11

.

In the shallow crust, magma degassing can also cause strong fractionation of D/H but

should produce a positive correlation with H2O due to the affinity of D for the vapor

phase25

. Positive trends are not, however, observed in the data, and we conclude that

degassing before entrapment did not play a significant role in modifying δD values of the

Cascade or Marianas samples.

Initial δD values (corrected as described above) from the Cascade melt inclusions

fall in a range from -95 to -70‰, despite a large range in H2O (Fig. 2.3a). These values

are isotopically lighter than those for the Marianas, which range from -55 to -12‰11

(Fig.

2.3a). Marianas melt inclusions have δD values that are higher than mantle values

(-80±10 ‰)26

, and have been interpreted to result from addition of D-enriched fluids to a

relatively D-depleted mantle wedge11

. Other estimates of δD for arc magmas based on

phenocrysts (-47 to -25‰ for Kamchatka)27

and fumarolic gases (-30 ± 10‰ for arcs

globally)28

agree well with the Marianas melt inclusion values. Although the Cascade

values overlap with mantle values, a minimum of 67–79% of the H2O in the Cascade

magmas must be sourced from a hydrous subduction component (Fig. 2.2a). We interpret

Page 31: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

15

the lighter isotopic values for the Cascades to represent fluids sourced from a plate that

has experienced more dehydration prior to reaching sub-arc depths than in the Marianas.

As the slab becomes progressively more dehydrated, it becomes isotopically lighter as D-

enriched fluids are removed. Thus, fluids released at higher temperatures and greater

extents of slab dehydration should have more negative δD values than fluids that were

removed from the slab at lower temperatures. An alternative possibility is that the

differences in δD between the Cascades and other arcs is the result of large-scale

diffusive fractionation of D and H in the slab, mantle and/or crust during migration of

fluids and melts. Although this cannot be ruled out, it would require that the transport of

H2O is largely diffusive rather than advective, which is inconsistent with field evidence

for fluid migration through fracture networks (see Supplementary Information A.).

2.4. D/H Fractionation Model

To test the interpretation that differences in δD between the Cascades and

Marianas are due to different extents of slab dehydration, we created a model that

describes D/H fractionation during dehydration reactions in the subducted plate beneath

the two arcs. Using an equilibrium fractionation equation (Supplementary Eq. 1), we

calculated the δD of fluids progressively released from the slab with increasing

temperature and depth. We calculate H2O released from the slab by coupling a 2-D

thermal model2 with the thermodynamic calculation code Perple_X

29,30. We divide the

slab into 1000-m wide vertical columns and calculate the cumulative δD value of fluid

Page 32: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

16

released from each column from the shallowest to the deepest column, assuming that the

slab consists of 2 km of basalt, 5 km of gabbro, and 2-4 km of hydrated mantle peridotite

(Fig. 2.3b, 2.3c; see Supplementary Fig. A.4 and Information), using temperature and

lithologically dependent fractionation factors (Supplementary Fig. A.5). A sedimentary

package was excluded from this model because most clay minerals break down at

shallow depths (5-10 km; ~250-450C) and contribute little to the fluid flux at sub-arc

depths6,21

.

Model results show that beneath the southern Cascades, the upper portions of the

slab, consisting of MORB volcanics and lower crustal gabbros, become completely

dehydrated by ~65 km depth (Fig. 2.3b), similar to previous predictions6. The hydrated

upper mantle portion of the subducted plate is able to carry bound H2O to greater depths

in the form of chlorite, but this phase is predicted to break down when the slab top

reaches ~80-90 km depth, which is directly beneath the volcanic arc (Fig. 2.3b). Our

model predicts two fluid pulses from the subducted plate: one beneath the forearc from

final dehydration of MORB volcanics and gabbros in the crustal part of the slab and

another beneath the arc from chlorite breakdown in the upper mantle portion of the slab.

The predicted depths of these fluid pulses correlate well with regions of low electrical

resistivity and low shear wave velocity in the overlying mantle wedge as imaged by

magnetotelluric and seismic data, respectively31,32

.

Page 33: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

17

Figure 2.3. Hydrogen isotopes: measured values and model results. a) H2O versus δD

for Lassen (green shaded region excludes inclusions severely affected by H-loss; colored

symbols are initial values; errors based on diffusion correction; Table 2.1; Supplementary

Information A.), the Marianas Arc11

, Koolau48

, Manus Basin49

, MORB26

and arc

fumaroles28

. b) Distribution of bound H2O in the slab calculated with our Lassen thermal

model (Supplementary Fig. A.4 and Information). Slab thickness exaggerated x3. c)

Fractionation model results. δD of fluids released from the entire slab package as a

function of slab top depth beneath the Cascades (green) and Marianas (blue). Uncertainty

is ±15‰ (shaded regions; 2σ; Supplementary Fig. A.4 and Information). Grey boxes

show range in melt inclusion data at sub-arc depths.

Page 34: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

18

At sub-arc depths beneath the Lassen region, modeled values of δD of the

hydrous component released from the slab overlap with corrected melt inclusion values

(Fig. 2.3b, 2.3c). This agreement suggests that Cascade arc magmas record the D/H ratio

of fluids derived from the deeper interior of the slab. The δD values of antigorite- and

chlorite-bearing peridotites, interpreted as subducted slab mantle, from Cerro del Almirez

(Spain)33

agree well with our model predictions (see Supplementary Information A.). In

contrast to the Cascades, model results for the Marianas (Fig. 2.3c), where older oceanic

crust is being subducted, show that temperatures within the slab remain cool enough to

carry bound H2O in hydrated upper mantle well past sub-arc depths (~160 km). In the

Marianas, modeled D/H values for fluids released from the plate overlap with the lowest

measured melt inclusion values. Although the modeled values do not reproduce the full

range of the data for the Marianas, the model does predict higher δD values for fluids

released beneath the Marianas arc compared to the Cascades.

2.5. Evidence and Implications for Slab Melting

Beneath the Lassen region, the upper, completely dehydrated portions of the plate

are likely at or above the MORB + H2O21

and sediment + H2O22

solidi (Fig. 2.2b). Our

model results predict that H2O-rich fluids are released from the slab interior during

chlorite breakdown at sub-arc depths, providing H2O that could cause partial melting of

the slab top. Although sediment melting imparts a distinctive trace element signature on

many arc magmas worldwide34

, in the Lassen Region it appears to be subordinate to

contributions from the subducted basaltic crust13

. Previous work in the Lassen

segment13,14

found that magmas with the highest (Sr/P)N, and therefore the greatest

Page 35: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

19

amount of subduction component, have the lowest Sr isotope ratio (87

Sr/86

Sr = 0.7029-

0.7034), similar to unaltered Gorda Ridge MORB35

(87

Sr/86

Sr = 0.7024; Supplementary

Fig. A.6). In light of our new results, we interpret the MORB-like 87

Sr/86

Sr signature to

reflect hydrous partial melting of Gorda MORB in which altered domains with radiogenic

Sr derived from seawater have lost much of their Sr during the shallower stages of

dehydration (see Supplementary Information A.). Strong negative correlation of 87

Sr/86

Sr

and (Sr/P)N in the Mt. Shasta region15

is also consistent with this interpretation, but

existing data for other segments of the arc further to the north do not show this pattern36,37

(Supplementary Fig. A.6).

Our results provide a mechanism for melting of sediments and basalt in the upper

part of the subducted oceanic plate by fluids rising from deeper parts of the slab17,38,39,40

.

Because melting of the upper oceanic crust takes place in the garnet stability field, a

hallmark of slab melting is the geochemical signature imposed by garnet, such as high

Sr/Y and La/Yb. Lavas from the Cascades and other relatively hot-slab subduction zones

like Mexico have values of these ratios that are higher than most other arc magmas, but

not as extreme as many so-called adakites, which are interpreted to be slab melts41

. Wet

melts of the basaltic crust are expected to be andesitic to dacitic in composition42,43

and

will infiltrate and react with the overlying mantle wedge and lower its solidus

temperature as they rise from the plate interface. A comparison of temperature sensitive

(H2O/Ce) and garnet sensitive (La/Yb) parameters for arc magmas worldwide

(Supplementary Fig. A.7) lends support to our interpretation that slab melting plays an

increasingly important role in arcs where younger, warmer oceanic crust is being

subducted.

Page 36: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

20

2.6. Bridge

In the preceding chapter (II), the hydrogen isotopes compositions of olivine-

hosted melt inclusions from seven cinder cones in the Lassen region were compared with

a thermo-petrologic model to investigate slab dehydration and melting beneath the

Cascade Arc. The next chapter (III) uses the volatile contents, major and trace elements,

and the B isotope compositions of olivine-hosted melt inclusions from eight cinder cones

(one in addition to the seven discusses in chapter II) to provide further evidence for the

conclusions drawn in chapter II and investigate magma generation processes in the

mantle wedge beneath the southern Cascades.

Page 37: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

21

CHAPTER III

SLAB MELTING AND MAGMA FORMATION BENEATH THE

SOUTHERN CASCADE ARC

This chapter includes material co-authored with Dr. P.J. Wallace, Dr. M.A.

Clynne, Dr. D. Weis, and Dan Rasmussen. Sample collection was performed by me with

the help of Dr. Wallace and Dr. Clynne. All melt inclusions were prepared and analyzed

by me using Fourier transform infrared spectroscopy (H2O and CO2), electron

microprobe (major elements), laser ablation inductively coupled plasma mass

Spectrometry (trace elements), and secondary ion mass spectrometry (B isotopes), with

assistance from D. Rasmussen for two cinder cones. Analyses, data corrections,

geochemical modeling, and interpretation were performed by me, with the help of Dr.

Wallace. Radiogenic isotopic analyses on bulk tephra were performed by Dr. Weis.

Models of mantle melting, using pMELTS, were performed by me. I drafted all the

figures used in this article and the text was written by me, with editorial assistance by Dr.

Wallace.

3.1. Introduction

Dehydration of subducted oceanic lithosphere drives arc magmatism at

convergent plate margins. However, the thermal structure of an individual subduction

zone controls the depths at which key dehydration reactions occur (Schmidt and Poli,

1998; Hacker et al., 2008; Van Keken et al., 2011; Wada et al., 2012), and consequently,

affects the composition and sources of fluids and/or melts (e.g., Cooper et al., 2012;

Page 38: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

22

Ruscitto et al, 2012) that infiltrate the mantle wedge to trigger melting. The thermal

structure is commonly assessed using the thermal parameter (Φ), which is a function of

downgoing plate age, dip angle, and convergence rate (Syracuse et al., 2010, Wada and

Wang, 2009). Variability in Φ globally is predicted to cause a wide range of slab surface

temperatures beneath arcs (give T range), as estimated from geodynamic models

(Syracuse et al, 2010; Wada and Wang, 2009) and geochemical tools (Plank et al., 2009;

Cooper et al., 2012). The results suggest a continuum of subduction zones between ‘cold’

(i.e., Tonga, Kamchatka) and ‘warm’ slabs (i.e., the Cascades, Mexico). Although fluids

released from the subducting slab have been shown to become more solute-rich with

increased temperature (Manning, 2004; Kessel et al., 2005a; Herman and Spandler, 2008;

Cooper et al, 2012; Ruscitto et al, 2012), whether or not the oceanic crust begins to melt

beneath arcs has been debated, despite geochemical evidence for slab melt contributions

beneath some warm-slab endmembers such Mexico (e.g., Cai et al., 2014), and the

Cascades (Walowski et al., 2015). Isolating the role of different components in the

subducted lithosphere (altered oceanic crust, sediments, serpentinized peridotite) and how

each component transfers material to the overlying mantle wedge (as fluids, melts or a

supercritical phase), is important for our understanding of mass transfer to the mantle,

both beneath and beyond arcs, as well as the processes by which the mantle melts in

subduction zones.

The Cascade arc represents a global warm-slab endmember due to slow shallow

subduction of young oceanic crust (6-10 Ma at the trench; Wilson et al., 2002; Kirby et

al., 2002). Estimations from geodynamic models (Syracuse et al., 2010; Wada and Wang,

2009) and geochemical studies (Cooper et al., 2012; Ruscitto et al., 2012; Walowski et

Page 39: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

23

al., 2015) agree that slab surface temperatures beneath the arc axis are hotter, on average,

than many other arcs globally. Previous work in the central Cascades has suggested that

the mantle wedge beneath the arc receives a reduced flux of volatiles from the downgoing

slab (Ruscitto et al., 2012), and H2O concentrations in olivine-hosted melt inclusions

from both the central and southern Cascades (~3.2 wt%; Ruscitto et al., 2010, 2011,

LeVoyer et al., 2010) fall slightly below the global average (~3.9 wt%; Plank et al.,

2013). Walowski et al. (2015) found that hydrogen isotope ratios of primitive magmas

from the Lassen region of the southern Cascades are lighter than those for the Marianas.

This was demonstrated to reflect waning dehydration of the deep slab interior (hydrated

mantle) after the crustal portion of the slab had already dehydrated beneath the forearc.

These results also provide evidence that flux-melting of the oceanic crust occurs when

the fluids released from the slab interior interact with oceanic crust that is above its wet

solidus temperature (i.e., Skora and Blundy, 2010; Till et al., 2012; Spandler and Pirard,

2013).

We measured the volatile contents, major element, trace element, and B isotope

compositions of olivine-hosted melt inclusions and the radiogenic isotopic compositions

of bulk tephra from the same eruptive centers in the Lassen region that were studied by

Walowski et al. (2015) for H isotopes. With these near-primary mantle melts, we aim to

quantify the chemical contributions from the subducting oceanic lithosphere to better

understand how subduction of warm oceanic crust affects the composition of mantle

melts and the productivity of melting in the mantle wedge. Lastly, we test the hypothesis

presented by Walowski et al. (2015) that magma production beneath the southern

Page 40: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

24

Cascade Arc involves a multi-stage process that includes flux melting of the subducted

oceanic crust and hydrous slab melt addition to the overlying mantle wedge.

3.2. Geologic Setting

The Lassen Segment is the southern terminus of the active Cascade arc (Guffanti

and Weaver, 1988). Volcanism is the result of oblique subduction of the Gorda micro-

plate beneath the North American plate (Fig. 3.1; Wilson, 2002), producing dominantly

calc-alkaline magmas (Clynne and Muffler, 2010). Westward expansion of the Basin and

Range extensional province into the eastern flanks of the Cascade Arc, including the Hat

Creek and Lake Almanor Grabens, has produced many normal faults that provide

pathways for mafic magmas to reach the surface (Guffanti et al., 1990; Clynne and

Muffler, 2010). The Quaternary volcanics in the Lassen region sit above a broad platform

of mafic to intermediate volcanoes and volcanic products 2-4 km thick (Berge and

Stauber, 1987). This volcanic platform is underlain by plutonic Sierran and metamorphic

Klamath terrain basement rocks as suggested by constant modeled seismic velocities

across the Sierra Nevada-Cascade Range boundary (Berge and Stauber, 1987).

Surrounding the volcanic center of Lassen Peak, a dacitic dome complex (Clynne

and Muffler, 2010), is a large volcanic field containing over 500 monogenetic cinder

cones and small shield volcanoes erupted in the last 12 Ma (Guffanti et al., 1990).

Previous work on these mafic cinder cones has identified a range in compositions from

low-K tholeiitic basalts to calc-alkaline basalts, basaltic andesites, and andesites (Clynne,

1993; Borg et al., 1997). Also observed in the most primitive calc-alkaline volcanic rocks

is distinct across-arc geochemical variability interpreted to be caused by variable

Page 41: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

25

Figure 3.1. Sample locations and across-arc geochemical variability. a) Regional map of

the Northwestern United States showing major tectonic boundaries. The Cascade

volcanic arc is defined by the major peaked labeled with black triangles. Lassen Peak is

highlighted in red. Black arrows show direction and are labeled with the convergence rate

relative to North America (Wilson et al., 2002). b) Larger scale map of the Lassen region

with locations of vents sampled in this study (BRVB: Basalt of Round Valley Butte,

BPB: Basalt of Poison Butte, BRM: Basalt of Red Mountain, BBL: Basalt of Big Lake,

BAS-44: Basalt of Hwy 44, BPPC: Basalt of Paine Parasitic Cone, BORG: Basalt of Old

Railroad Grade, and CC: Cinder Cone; see Table 3.1 for details) and previously sampled

by Clynne (1993) and Borg (1995; grey diamonds). Lassen Peak (large white triangle),

outcropping basement rocks (shaded pink areas), major highways (thin black lines), and

large lakes (shaded blue regions), are also highlighted. c) Distance from the trench vs.

Sr/Nd and d) 87Sr/86Sr for samples in this study(colored symbols) and Borg et al. (1997;

grey diamonds). Symbols and colors for individual cinder cones are consistent throughout

the manuscript.

Page 42: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

26

enrichment of the mantle by a subduction component, a key factor for targeting this

region in our study (Fig. 3.1; Borg 1997, 2002). Figure 3.1 shows variations in both

Sr/Nd and 87

Sr/86

Sr with increasing distance from the trench. Decreasing Sr/Nd has been

interpreted to indicate the waning addition of a subduction component towards the back-

arc (Borg et al., 1997). The 87

Sr/86

Sr compositions, however, increase toward the back-

arc, indicating that the subduction component has a less radiogenic Sr isotope signature

than the sub-arc mantle. Furthermore, variability in Nb/Zr suggests that the Lassen sub-

arc mantle is heterogeneous (Borg et al., 1997; Walowski et al., 2015), but there are no

systematic variations of Nb/Zr with distance from the trench.

3.3. Sample Descriptions and Analytical Methods

Samples were collected from the tephra deposits of Quaternary monogenetic vents

spanning ~80 km from the forearc to the backarc (Fig. 3.1). Vents that erupted primitive

basalt or basaltic andesite (MgO > 7 wt%) identified from bulk rock analyses (Clynne,

1993; Borg et al., 1997) were targeted because they are close to primary mantle melts in

composition. Coarse ash was collected because it quenches more rapidly than lava,

bombs, and lapilli, minimizing the potential for syn-eruptive diffusive H loss or

crystallization of melt inclusions (e.g., Lloyd, 2013). Loose olivine crystals (50 μm to 1

mm in size) were hand-picked from washed and sieved tephra. The olivine crystals were

treated in HBF4 to remove glass adhering to the crystal surface and examined in

immersion oil (refractive index 1.675) to locate melt inclusions. Olivine crystals hosting

fully enclosed, glassy melt inclusions were mounted in acetone-soluble resin on glass

slides and prepared as doubly polished wafers. H2O and CO2 concentrations of the melt

Page 43: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

27

inclusions were measured at the University of Oregon using a Thermo-Nicolet Nexus 670

FTIR spectrometer interfaced with a Continuum IR microscope. Concentrations were

calculated from IR peak absorbances using the Beer–Lambert law and compositionally

appropriate absorption coefficients (see Johnson et al., 2008, for details). Melt inclusions

and host olivine were analyzed for major elements (plus S and Cl for inclusions) on the

Cameca SX-100 electron microprobe at the University of Oregon (see Ruscitto et al.,

2010, for details). Melt inclusions were subsequently analyzed for a suite of trace

elements on the Photon Machines Analyte G2 193 nm ArF ‘‘fast’’Excimer Laser system

at Oregon State University, using 50 µm spot size with a 5 Hz pulse rate. Measured trace

element concentrations were determined by reference to GSE-1G glass as a calibration

standard and using 43

Ca as an internal standard (see Loewen and Kent, 2012, for details).

BHVO-2G, BCR-2G, and GSD-1G glasses were also analyzed to monitor accuracy and

precision, and the analyzed values were within 10% of accepted values.

A subset of the melt inclusions that were analyzed for H isotopes and trace

elements by Walowski et al. (2015) were also analyzed for B isotope ratios using the

Cameca IMS 1280 at the Northeastern National Ion Microprobe Facility, Woods Hole

Oceanographic Institution, with O- primary beam, 20-40 nA primary current, 10,000 V

secondary voltage, and a 10 µm spot size. More detailed methods are described in

Marschall and Monteleone (2014). Some of the melt inclusions were too small to allow a

new SIMS spot adjacent to an existing NanoSIMS spot (20x20 rastered area, ~5 µm

deep). In these cases, the SIMS spot was placed within the pre-existing NanoSIMS spot.

Tests comparing measurements within pre-existing spots to those on a clean surface from

a single melt inclusion revealed no systematic differences.

Page 44: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

28

Sr, Nd, Hf, and Pb isotope ratios were measured at the Pacific Centre for Isotopic

and Geochemical Research at the University of British Columbia. Pb, Nd, and Hf isotope

ratios were measured by MC-ICP-MS (Nu Instruments Ltd., Nu II 214), and Sr isotope

ratios were measured by Thermo Finnigan Triton TIMS using procedures described in

Weis et al. (2007) and Weis et al. (2006). Additional details regarding sample preparation

and analytical techniques are given by Mullen and Weis (2015). Tephra from sample CC

was excluded from isotopic analyses due to clear crustal contamination (abundant quartz

xenocrysts) of the bulk material.

3.4. Results

3.4.1. Melt Inclusion Major and Trace Element Compositions

The olivine host crystals vary from Fo84 to Fo90 (Table 3.1). For each cinder cone,

11-17 melt inclusions were analyzed and individually corrected for post-entrapment

crystallization (PEC) and Fe-loss using Petrolog 3.1 (Danyushevsky and Plechov, 2011;

Danyushevsky, 2000). Volatile contents and trace elements, which are not included in the

Petrolog calculations, were also corrected using the ratios of K2O in corrected and

uncorrected inclusions. Initial Fe contents were based on the FeOT of the bulk tephra or

individual melt inclusions with the highest MgO and FeOT at a single cone (bulk tephra

compositions were not available for cones BBL and CC). Calculated percentages of PEC

for all melt inclusions range from 0 - 14%. Corrected melt inclusion

compositions overlap with the most primitive lava samples previously analyzed in the

Lassen region (Fig. 3.2) and contain MgO concentrations of 7.4 - 9.8 wt%

Page 45: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

29

Table 3.1. Averaged Primary Melt Compositions.

Page 46: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

30

Figure 3.2. Corrected melt inclusion major and trace element compositions. a) Average

melt inclusion trace element composition for each vent normalized to normal-MORB (N-

MORB; Sun and McDonough, 1989) and compared with previously determined end-

member compositions (CAB, Borg et al. 1997; HAOT, Bacon et al. 1997). b) K2O and

SiO2 contents of individual melt inclusions (corrected, volatile-free normalized values)

compared with bulk rock analyses from Clynne, (1993; grey diamonds). Classification of

low-, med-, and high-K based on Peccerillo and Taylor (1976).

(Supplementary Table B.1; see Apendix B for all supplementary information for this

chapter). To estimate a primary melt composition for each cone, we incrementally added

equilibrium olivine (0.1 wt% steps) to the average melt inclusion composition from each

cone until the melt composition was in equilibrium with mantle olivine (Fo90; Table 3.1;

Page 47: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

31

Ruscitto et al., 2010). The calculated primary melt compositions required addition of 1-

20% equilibrium olivine (Table 3.1).

The melt inclusion glasses are dominantly medium-K calc-alkaline basalt (CAB),

with some that fall into the low-K field (Fig. 3.2a). Previous work in the Lassen region

has suggested that low-K tholeiitic magmas (also called high-alumina olivine tholeiites;

HAOT) and calc-alkaline magmas have fundamentally different source regions (Clynne,

1993; Bacon et al., 1997; Baker et al., 1994). However, the low-K samples used in this

study do not display the lower LREE/HREE and LILE/HFSE values typical of the

endmember HAOT volcanic rocks in this region (Fig. 3.2, Bacon et al., 1997). Rather, all

compositions used in this study display similar trace element patterns to the regional

CABs (Fig. 3.2), suggesting that despite variability in major and trace element

compositions, they likely formed as the result of flux-melting, and not decompression

melting of the mantle.

3.4.2. Magmatic Volatile Contents

Dissolved H2O contents of the melt inclusions range from 0.6-3.5 wt%. At

individual cinder cones, a range in H2O concentrations is observed and is likely due to

differences in extent of pre-entrapment degassing (Metrich and Wallace 2008; Johnson et

al., 2009) and post-entrapment hydrogen loss (Lloyd et al., 2013, Gaetani et al., 2012,

Bucholz et al., 2013). Because these processes systematically decrease H2O

concentrations, the maximum measured H2O/K2O ratio for each cone was used to

estimate the initial H2O content (H2Omax) of the magma erupted at that cone. In the

Lassen region, H2Omax ranges from 1.3-3.4 wt%. Primary melt H2O concentrations (1.1-

Page 48: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

32

3.4 wt%, calculated from H2Omax using the olivine addition method described above;

Table 3.1) overlap with calculated primary melt H2O values for basaltic and basaltic

andesite melt inclusions from central Oregon (1.4-3.0 wt%; Ruscitto et al., 2010). In

contrast to H2O, Cl is not affected by degassing and post-entrapment diffusive effects.

Concentrations of Cl in calculated primary melts range from 100-600 ppm, except at

BRM, where Cl values are as high as 2500 ppm (Supplementary Table B1).

3.4.3. Isotopic Compositions

The δ11B values from individual cinder cones in the Lassen region range from

-9.9‰ to -2.4‰ (Table 3.1; Fig. 3.3). These values overlap with those measured for bulk

rock samples from the southern Washington Cascades, which range from -9‰ to -0.4‰

(Leeman et al., 2004), and those from melt inclusions from Mt. Shasta (Fig. 3.3; Rose et

al., 2001; LeVoyer et al., 2010). Melt inclusions from the Cascades have lower B

concentrations and more negative δ11B than those measured in other arcs, such as

Kamchatka and Japan, where older oceanic crust subducts (Fig. 3.3; Ishikawa et al.,

2001; Ishikawa and Tera, 1999).

The Sr, Nd, Hf, and Pb isotope measurements for bulk tephra samples overlap

with those previously determined for volcanic rocks in the Lassen Region (Borg et al.,

1997; Borg et al., 2002) and show variability in 87

Sr/86

Sr (0.703396-0.703985±0.000007),

208Pb/

204Pb(38.4640-38.6503±0.0026),

177Hf/

176Hf (0.283035-0.283094±0.000006), and

143Nd/

144Nd (0.512827-0.512948±0.000006) (Table 3.1).

Page 49: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

33

Figure 3.3. Boron isotope compositions. Each data point (symbols as in Fig. 3.1)

represents an average from 4-6 individual melt inclusion analyses. Data for comparison

from the southern Washington Cascades (Leeman et al., 2004; filled squares, whole rock

analyses) Mt. Shasta (Rose et al., 2001, open circles; individual melt inclusions; LeVoyer

et al., 2010, filled circles), the Marianas (Ishikawa, 2001), and Kamchatka (Ishikawa and

Tera, 1999). Dashed black curve represent mixing between MORB (Chaussidon and

Jambon, 1993) and a hydrous slab fluid (Marschall, 2007).

3.5. Discussion

3.5.1. The Source of Volatiles in Lassen Region Primitive Magmas

B is a highly fluid mobile element, and because it is found in higher

concentrations in subducted materials than the mantle, it can be an excellent tracer of

fluids derived from subducting slabs (e.g., Bebout et a.,1993, Ryan and Langmuir 1993,

Tonarini et al., 2001). In addition, subducted materials such as sediment, oceanic crust,

and serpentinitized mantle have δ11B signatures that are distinct from the mantle wedge

(e.g., Ishikawa and Nakamura, 1993; Smith et al., 1995; Boschi et al., 2008; Chaussidon

and Jambon, 1994). However, the B isotopic composition and low B concentrations in the

Page 50: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

34

melt inclusions from the Lassen region suggest that the sub-arc mantle receives little to

no B from the subducting slab (Fig. 3.3). This is probably the result of extensive

dehydration of the slab before it reaches sub-arc depths (Leeman et al., 2004; Manea et

al., 2014). Geodynamic modeling and calculated metamorphic phase equilibria suggest

that H2O can be carried to sub-arc depths by the hydrated mantle portion of the

downgoing slab, in the form of chlorite, beneath the Lassen region (van Keken et al.,

2011; Walowski et al., 2015). The D/H ratios measured in melt inclusions from the

Lassen region are in agreement with modeled values which indicate that fluids are likely

sourced from the final breakdown of chlorite in the deep slab interior (Walowski et al.,

2015). Because nearly all B is released from hydrated peridotite during the antigorite

breakdown reaction (which occurs in the slab beneath the Cascades before reaching sub-

arc depths), chlorite-derived fluids contribute little to no B to the subduction component

(Spandler et al., 2014). This explains how the slab beneath the Cascades can release a

hydrous slab component to the mantle wedge that contains very little B, such that

primitive magmas formed in the wedge have B isotope ratios and concentrations typical

of mantle-derived magmas (MORB).

Despite low B concentrations, H2O and Cl are high compared to MORB, which

requires that these volatiles are retained in the slab to greater depths than B (Spandler et

al., 2014). Furthermore, strong correlations of H2Omax/Ce and Cl/Nb with Sr/Nd clearly

demonstrate that volatile and trace element enrichments are coupled and therefore are

derived from the same process (Fig. 3.4). To better quantify this relationship, we

calculated partial melt compositions for two mantle endmembers to which variable

amounts of subduction component were added. Figure 3.4 shows good agreement

Page 51: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

35

between the model curves and the melt inclusion data, which indicates that volatile and

trace element variability at different cones is the result of different amounts of a

subduction component added to a heterogeneous mantle wedge. However, in some cases,

such as BRM, a sample will have a lower H2O/Ce for a given value of Sr/Nd than

predicted by the melting model. This may be caused by variability in the H2O and trace

element ratios of the hydrous subduction component, or it may suggest that at some

cones, H2Omax is still a partly degassed composition (either because of pre-entrapment

degassing or post-entrapment H loss). Cl/Nb should provide a more robust indication of

initial volatile concentration because Cl is not affected by diffusive loss and only

degasses at very low pressure. Good agreement between the data and melting models for

Cl/Nb vs. Sr/Nd provides support for the interpretation that initial H2O concentrations are

related to the amount of a subduction component added to the mantle wedge beneath the

arc and that the slab component has ratios of H2O and Cl to LILE that are not highly

variable (Fig. 3.4b and c). This evidence indicates that BRM, the sample with the highest

Sr/Nd and therefore largest amount of a subduction component, has very low H2O/Ce as

the result of extensive degassing in the crust or post-entrapment hydrogen diffusive loss.

Volatile and trace element ratios from the central Oregon Cascades can also be

explained using the calculated melting curves, but require subduction component addition

to a more enriched mantle source than most Lassen region magmas (Fig. 3.4a; Ruscitto et

al., 2010). Interestingly, melt inclusions with the highest values of Sr/Nd in both the

Lassen region (BRM) and the Mt. Shasta region do not have the highest values of

H2O/Ce, but they do have the highest Cl/Nb, and also have Cl concentrations

significantly higher than other cones throughout the Cascades. As suggested above, these

Page 52: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

36

magmas likely experienced extensive degassing of H2O in the crust before melt inclusion

entrapment, and/or were affected by post-entrapment H loss. However, if the BRM and

Mt. Shasta magmas had pre-degassing compositions that fit the model curves in Figure

3.4a, they would have initially had up to ~8-10 wt% H2O. This is consistent with

experimental phase equilibria showing that high-Mg basaltic andesites and andesites at

Mt. Shasta had very high initial H2O contents (10-14 wt%; Krawczynski et al., 2012).

Such high initial H2O is also consistent with the very high Cl/Nb values for high-Mg

basalts and basaltic andesites at Mt. Shasta (Ruscitto et al., 2011).

Figure 3.4. Corrected melt inclusion volatile and trace element ratios. a) H2O/Ce vs.

Sr/Nd (corrected melt inclusions that contain H2O concentrations within 0.5 wt% of the

H2Omax from each cone to represent the least degassed compositions), b) Cl/Nb vs. Sr/Nd

(all melt inclusions; corrected) and c) Cl/Nb vs. Sr/Nd (y-axis extents to higher values,

average values calculated using corrected melt inclusions from each cone). Data from

central Oregon (Ruscitto et al., 2010; restored melt inclusions; solid blue circles enclosed

in light blue shaded field) and Mt. Shasta (Ruscitto et al., 2008; restored melt inclusions;

primitive basaltic andesite (PBA): solid grey triangles enclosed in grey shaded field;

high-Mg andesites (HMA): open grey triangles enclosed in a grey shaded field), for

comparison. Black lines represent 10% partial melts of two endmember mantle

compositions (DMM; Workman and Hart, 2005; average central Oregon mantle; Ruscitto

et al., 2010) mixed with variable amounts of a hydrous subduction component (grey

diamond; calculated using methods of Portnyagin et al. (2007), with primary mantle

composition of BORG; Table 3.1). Grey bar represents the range in sub-arc mantle

compositions determined by Walowski et al. (2015). Melt inclusions that have

experienced degassing or syn-eruptive H loss will cause deviation from the melting

curves as indicated by the black arrow in panel (a).

Page 53: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

37

The Blanco Fracture zone, which separates the Juan de Fuca and Gorda plates,

may provide a pathway for deep serpentinization of the upper mantle in the downgoing

slab offshore of the Cascades, and has been proposed to source the volatile-rich

component at Mt. Shasta (Grove et al., 2002; Manea et al., 2014) However, plate

reconstructions and symmetry of the fracture zone on the Pacific plate suggest that the

Blanco Fracture zone is not old enough to project beneath the arc (Wilson, 2002), and

thus, the causes of the geochemical differences between the Mt. Shasta and Lassen

regions (Fig. 3.4) remain enigmatic.

3.5.2. Evidence for Slab Melting

Previous work using trace elements and radiogenic isotopes in the Lassen region

found negative correlations between LILE/LREE ratios and 87

Sr/86

Sr, which suggests that

the subduction component is less radiogenic than the sub-arc mantle (Borg et al., 1997,

2002). New radiogenic isotope data from this study overlap with the published data (Fig.

3.5). Because elevated H2O/Ce, Cl/Nb, and Sr/Nd ratios are related to subduction

component addition, our data confirm that the subduction component is MORB-like, with

less radiogenic Sr and Pb isotopic composition and more radiogenic Nd isotope

composition than the Lassen region sub-arc mantle. In its isotopic characteristics, the

subduction component beneath the Lassen region is similar to offshore Gorda Ridge

MORB (Davis et al., 2008). However, a major contribution from the melting of

subducted MORB crust was originally discounted because dry eclogitized oceanic

lithosphere requires higher temperatures than expected for the slab top at sub-arc depths

(Borg et al., 1997). Similar isotopic and trace element relationships are observed in the

Page 54: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

38

Mt. Shasta region (Fig. 3.5). Grove et al. (2002) discounted slab melting beneath Shasta

as a possible explanation because models of hydrous peridotite melting could reproduce

the observed major element compositions of primitive volcanic rocks in that region.

Recent work by Walowski et al. (2015) interpreted the light D/H values of melt

inclusions from the Lassen region as resulting from final dehydration of chlorite in the

hydrated upper mantle portion of the downgoing slab. This provides a mechanism to

deliver H2O to the MORB-volcanic portion of the slab top and drive wet slab melting

beneath the arc.

To further test this hypothesis, we calculated mixing and partial melting models

involving the Lassen sub-arc mantle and a partial melt of Gorda Ridge MORB (Fig. 3.5).

Because temperatures of the plate top are at or above the wet MORB and wet sediment

solidi (Schmidt and Poli, 1998; Herman and Spandler, 2008), we assume the subduction

components are partial melts rather than aqueous fluids (Cooper et al., 2012; Ruscitto et

al., 2012; Walowski et al., 2015). We also assume that the Gorda MORB component is

unaltered. This assumption relies on the fact that oceanic crust is heterogeneously altered

during hydrothermal alteration (e.g., Bach, 2003), and that the unaltered portions retain

their MORB-like signature, whereas altered zones likely lose most of their seawater-

enriched Sr beneath the forearc during transition to ecologite (Walowski et al., 2015).

Because Sr/Nd is an indicator of subduction enrichment, primitive basalts and basaltic

andesite with the lowest Sr/Nd values should be most representative of the Lassen sub-

arc mantle. For the samples with the lowest Sr/Nd, a small range of Sr, Nd, and Pb

isotope ratios are observed, and these ranges are probably indicative of mantle

heterogeneity beneath the arc, as previous suggested by Borg et al. (1997; 2002) and

Page 55: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

39

Figure 3.5. Radiogenic isotope compositions and slab melt mixing models. a) Bulk

tephra compositions of a) 87

Sr/86

Sr b) 208

Pb/ 204

Pb, and c) 143

Nd/144

Nd vs. average Sr/Nd

from melt inclusions for each cone (symbols as in Figure 3.4). Isotopic composition of

BRM is from Borg et al. (1997), as it was not analyzed in this study. CC was omitted due

to evidence for crustal contamination (see main text). Bulk lava analyses from the Lassen

Region and M. Shasta (Borg et al., 1997; Grove et al., 2002, respectively) for

comparison. North Cascade sediment (yellow shaded region; Carpentier et al., 2013,

2014) and northern Sierran granites (grey shaded region; Cecil et al., 2012) highlighted to

show possible components that may contribute to trace element and radiogenic isotope

variability of samples. Melting models (dashed lines) calculated using the batch melting

equation for a range in mantle sources (calculated for the Lassen sub-arc mantle; see text)

mixed with 2, 5, and 10% (labeled on modeled curves) of a 5% partial melt of Gorda

MORB (Davis et al., 2008; partition coefficients from Kessel et al., 2005a; at 4 GPa,

1000C). Bulk partition coefficients calculated for a spinel peridotite assemblage

35/30/12/5-Ol/OPX/CPX/Sp (Hughes and Taylor, 1986) using mineral partition

coefficients of Eiler et al. (2005) for Sr and Nd. Melt fractions derived from PMELTS

model results (for a given temperature and amount of slab melt addition; Fig. 3.6). d)

Lassen sub-arc mantle (based on bulk rock sample with smallest amount of apparent

subduction component) mixed with sediment partial melts (5%, large filled grey

diamond, and 20%, large filled grey diamond, melts of N. Cascade sediment using

partition coefficients from Kessel et al. (2005) a) and slab partial melts 5%, large filled

black diamond, and 20%, large filled black diamond, melts of Gorda MORB using

partition coefficients from Kessel et al. (2005a).

Page 56: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

40

Clynne and Borg (1997). We thus use a range in sub-arc mantle compositions (Fig. 3.5;

87Sr/

/86Sr = 0.7039 - 0.7043,

208 Pb/

204Pb = 38.512 - 38.782 , and

143 Nd/

144Nd = 0.5127-

0.5128). Figure 3.5 shows curves that represent mantle melts formed by melting of the

sub-arc mantle mixed with variable amounts of Gorda MORB melt. The model results

suggests 1-10 wt% addition of a subduction component that is a slab melt. Although the

model results can explain a majority of the compositions, they under-predict some values

of Sr and Pb isotopes, and overpredict the range in Nd isotopes. There are three potential

explanations for these small offsets: 1) zones of alteration in the MORB volcanic section

of the downgoing plate that retained their isotopic signature even after complete

dehydration beneath the forearc, 2) involvement of small proportions of a sediment melt

component, and/or 3) contamination by continental crust. All three components may play

a role in causing the slight offset of the slab melt models from the Lassen data seen in

Figure 3.5. The latter two explanations will be discussed in further detail in the following

section.

Our proposed mechanism for slab melting relies on breakdown of chlorite in the

mantle portion of the downgoing oceanic plate. However, in the model shown in Figure

3.5, this component has been neglected in our melting calculations. Fluids produced from

the breakdown of chlorite at sub-arc depths have some distinct trace element

characteristics (e.g., elevated LREE/HREE), but overall, are solute poor (Spandler et al.,

2014), and as a result, will have little effect on the trace element composition of the

resulting magmas formed by flux melting of the upper oceanic crust. For example, Sr

concentrations in fluids produced from the breakdown of serpentine and chlorite range

from 1.7-2.8 ppm (Spandler et al., 2014), an insignificant contribution when compared to

Page 57: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

41

the average Sr composition of oceanic crust, 90 ppm (Sun and McDonough, 1989). We

therefore conclude that fluids derived from chlorite breakdown in the hydrated mantle

portion of the slab dominantly contribute H2O to the system but do not impart a

distinctive trace element signature.

3.5.3. The Role of Sediment Melts and Crustal Assimilation

Low Sr/Nd magmas in the Lassen region have elevated 207

Pb/ 204

Pb, which was

previously used to suggest the sub-arc mantle was enriched by a sediment component, but

it was concluded that this took place during an earlier, possibly Mesozoic, subduction

event because young Pacific sediments did not have high enough Pb isotope ratios to

explain the values (Borg et al., 1997). However, radiogenic isotope ratios measured in

sediments offshore of Vancouver Island (Carpentier, 2014; Mullen and Weis, 2015) do

have high enough Pb isotopic ratios to explain the isotopic compositions of low Sr/Nd

magmas in the Lassen region. This new data allows for an alternative hypothesis that

variable proportions of hydrous melts of subducted oceanic crust and sediment make up

the subduction component in the forearc, whereas sediment melts dominate in the back-

arc. This hypothesis, however, requires bulk sediment addition to the mantle wedge

rather than sediment melting (Fig. 3.5). Because sediment melts are more physically

plausible given the relatively high slab top temperatures beneath the arc, this component

should display higher Sr/Nd than bulk sediment, similar to MORB partial melts, but does

not (Fig. 3.5). Furthermore, because thermo-petrologic models suggest that the slab

becomes completely dehydrated at sub-arc depths, this would require in-situ breakdown

of phengite at great depths to drive sediment melting behind the arc-axis (Spandler and

Page 58: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

42

Pirard, 2013). However, low-velocity anomalies agree well with release of fluids from

final chlorite breakdown in the mantle portion of the slab at 80 km, and no further

evidence provides support for contributions from the slab past this depth. These

observations taken together suggest that subduction of a modern sediment component is

not responsible for the isotopic compositions of low-Sr/Nd magmas in the Lassen region

(Fig. 3.5).

Conversely, a modern sediment component may contribute to variability in

isotopic compositions of high-Sr/Nd magmas that trend to slightly more radiogenic

values of Sr and Pb, and less radiogenic values of Nd than predicted by the melting

model in Figure 3.5. To further distinguish between sediment and slab melt contributions,

we use Th/La systematics as a discriminant because of the high Th concentrations found

in sediments relative to MORB and sub-arc mantle (Fig. 3.6; Plank et al., 2005). The

mixing model in Figure 3.5d shows that Th/La variations in primitive Lassen region

magmas may result from addition of <5% of a subduction component that is made up of

variable proportions of sediment and MORB melts. At all but two cinder cones, melt

inclusion values suggest that <30% sediment melt contributes to the trace element and

isotopic enrichments.

Two cinder cones (BPB and CC) have larger apparent contributions from

sediment melts. Cinder Cone in particular contains the highest Th/La values from our

dataset. However, the bulk lava and tephra at this cone contain abundant quartz

xenocrysts and variably melted granitic xenoliths, which are clear indications of large

amounts of crustal contamination (M. Clynne, unpublished data). This sample was

therefore excluded from radiogenic isotope analyses. We also note that in general, melt

Page 59: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

43

inclusions from the Lassen region samples have lower Th/La values on average than

many of the published bulk rock analyses from this area (Fig. 3.5; Borg et al., 1997,

2002). This may be because melt inclusions are trapped at depth, before even minor

crustal contamination occurs in the shallow crust prior to eruption. Previous work in the

Lassen region interpreted unradiogenic Pb and Os isotopic compositions in the relatively

primitive basalts and basaltic andesites as evidence for minimal contamination by

continental crust (Borg et al., 1997, 2000). However, because of very high Th

concentrations in the granitic basement rocks (Cecil et al., 2012), small amounts of

crustal addition could increase the Th/La ratio enough to make it difficult to differentiate

between sediment melt contribution and that of the basement rocks.

3.5.4. Modeling Slab Melt Addition to the Mantle Wedge

To determine the influence of slab melt addition to the mantle wedge, we used

pMELTS (Ghiorso et al., 2002) to compare the effects of fluid addition and hydrous melt

addition to the mantle wedge at temperatures and pressures expected for the Lassen sub-

arc mantle. To do this, we created a mantle-wedge source composition by adding various

amounts of either a dacitic slab melt (Klimm et al., 2008) or pure H2O to a primitive

mantle composition (MM3; Baker and Stolper, 1994). The pMELTS program was then

used to determine the phase equilibria of the bulk mixture from 900-1400 C at a pressure

of 1.5 GPa. These values are based on temperatures from geodynamic model results for

the Lassen region (Walowski et al., 2015), beginning at the slab-wedge interface to

~100°C hotter than peak temperatures expected in the mantle wedge. Although this

method does not attempt to model the kind of reactive transport mechanism that likely

Page 60: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

44

occurs in the mantle wedge, it is conceptually similar to the approach used by

experimental petrologists to simulate mantle melting (e.g., Grove et al, 2002), which

involve equilibration of a given bulk composition at various temperatures and pressures.

Figure 3.6 compares the fraction of melt produced at various temperatures for a

given starting H2O concentrations in the model mantle sources generated by addition of

fluid or melt. Melt fractions for both the hydrous-melt-fluxed and fluid-fluxed peridotite

cases are nearly indistinguishable. This suggests that the amount of H2O supplied to the

mantle strongly controls the degree of melting irrespective of whether the H2O is added

as a hydrous melt or a fluid. Slightly higher melt fractions are seen in the melt-fluxed

melting case, which may result from the additional fluxing effect of alkalis on the

peridotite solidus (Hirschman et al., 1998).

Figure 3.6. Percent melt and source H2O: results from pMELTS. Relationships between

the degree of partial mantle melting and amount of wt% H2O in the starting mantle

component at a range of temperatures calculate with pMELTS (Ghiorso et al., 2002). The

H2O in the mantle source (MM3; Baker and Stolper, 1994) was varied by addition of pure

H2O (black dashed lines) and hydrous eclogite partial melts (solid black lines; dacite

major element composition with 8 wt% H2O; Klimm et al., 2008).

Page 61: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

45

Figure 3.7 shows the major element compositions of partial melts resulting from

various amounts of slab melt and aqueous fluid addition to the mantle wedge. For small

amounts of slab melt addition (1-5 wt%), the major element compositions of resulting

basaltic melts are similar to those of the aqueous fluid addition case. This shows that

equilibrium between partial melt and residual mantle largely controls the major element

composition of the final melt. Primary magma compositions calculated from the melt

inclusion data have compositions that overlap with the pMELTS model results (Fig. 3.7),

demonstrating that hydrous-melt-fluxed melting of the mantle wedge is a viable

explanation for the production of these magmas. However, when slab melt addition

exceeds 10%, the resulting high bulk H2O and alkali contents of the fluxed mantle wedge

source lead to higher melt fractions at a given temperature, and thus, exhaustion of

clinopyroxene occurs at ~1200C, producing melts with higher SiO2 at peak mantle

wedge temperatures. These high-Mg basaltic andesites may be analogous to natural high-

Mg basaltic andesite seen in the Lassen and Mt. Shasta regions (e.g., M. Clynne unpub.

data; Grove et al., 2002; Ruscitto et al., 2011).

Figure 3.8b displays the pMELTS model curves for Y and H2O for various

amounts of slab melt addition and temperatures. The H2Omax compositions and average Y

concentrations from the Lassen melt inclusions overlap with the model results and are

consistent with peak mantle wedge temperatures of 1300-1375C. The excellent

agreement between measured and modeled results, particularly with respect to H2O,

further supports the interpretation that the Lassen primitive magmas were formed by

variable addition of a hydrous slab melt component at similar peak mantle wedge

temperatures. Values of peak mantle temperatures overlap with those from Till et al.

Page 62: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

46

(2012), which predicts temperatures of last mantle equilibration of ~1250-1350C.

Slightly higher temperatures determined in our study are likely the result of known

temperature over-predictions made by pMELTS of ~50 C when compared with

experimental methods (Ghiorso et al., 2002).

Figure 3.7. Comparison of pMELTS model results and melt inclusion compositions (in

eq. with Fo90 olivine; Table 2). a) H2Omax (calculated primary composition), b) K2O +

Na2O, c) CaO and d) Al2O3 wt% vs. SiO2 (all major elements normalized on a volatile-

free basis). Phase equilibria were calculated using pMELTS for a mantle source (MM3;

Baker and Stolper, 1994) mixed with either pure H2O (dashed lines) or a hydrous dacite

melt (solid lines; Klimm et al., 2008). Each line represent a model result from 900-

1400C at 1.5 GPa (normalized on a volatile-free basis).

Page 63: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

47

3.5.5. The Sr/Y Adakite Signature

One hallmark characteristic of slab melt in arc magmas is high Sr/Y, caused by

melting in the presence of garnet, which makes up a large proportion of ecologitized

MORB in the downgoing oceanic lithosphere (Duffant and Drummond, 1990). During

ecologite melting, Sr is incompatible and increases in the melt phase, whereas Y is

compatible with garnet and retained in the downgoing slab. Most Lassen Region

magmas, however, do not display high Sr/Y values compared with the global array of

adakitic magmas (Fig. 3.8b). Using the same mixed mantle compositions and melt

fractions derived from the pMELTS models, melting curves in Figure 3.9b show that for

small amounts of slab melt addition (1-10%, similar to major element results), the Sr/Y

ratio is dampened due to the addition of Y from the spinel peridotite mantle component.

This yields values that overlap with values measured in melt inclusions from all but one

sample (BRM) from the Lassen segment (Fig. 3.8b). The results are consistent with

previous calculations from Kelemen (1993), which showed that peridotite-melt reactions

produced melts with lower LREE/HREE than the initial slab melts. The pMELTS models

results indicate that high Sr/Y adakitic signatures may only be retained in arc magmas if

slab melt addition is >10 wt%. Larger proportions of slab melt addition are thus required

to explain the high Sr/Y value of BRM, consistent with estimates of ~10% slab melt

addition derived from radiogenic isotope melting models (Fig. 3.5), and predictions that

this sample may have initially contained much higher H2O concentrations (Fig. 3.4). The

high-Mg andesites from the Lassen Region (M. Clynne, unpub. data) and the Mt. Shasta

region that have higher values of Sr/Y (~150; Ruscitto et al., 2011) could therefore be

produced by larger amounts of hydrous slab melt addition to the mantle wedge.

Page 64: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

48

Figure 3.8. The effect of slab melt addition to peridotite for Sr/Y, Y and H2O. a) H2Omax

and b) Sr/Y with average Y for each cone (calculated primary compositions; Fo90; Table

3.1) with global range of adakites (GEOROC database), and experimental eclogite partial

melts (Klimm et at., 2008). Solid and dashed curves represent modeled mantle melt

compositions for various amounts of slab melt addition from 900-1300C at 1.5 GPa.

Modeled H2O composition derived as in Fig. 3.7. Modeled Sr and Y were calculated

using the batch melting equation for a mantle source (calculated for the Lassen sub-arc

mantle) mixed with 2, 5, and 10% (labeled on modeled curves) of a 5% partial melt of

Gorda MORB (Davis et al., 2008; partition coefficients from Kessel et al.,2005a; at 4

GPa, 1000C). Bulk partition coefficients calculated for a spinel peridotite assemblage

35/30/12/5-Ol/OPX/CPX/Sp (Hughes and Taylor, 1986) using mineral partition

coefficients of Eiler et al.(2005) for Sr and Eiler et al. ( 2001) for Y. Melt fractions

derived from pMELTS model results (for a given temperature and amount of slab melt

addition; Fig. 3.6).

Page 65: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

49

Although most primitive Cascade arc magmas do not have particularly high Sr/Y

values compared to adakites, they do have other characteristics that indicate melting in

the presence of garnet when compared to the global array of basaltic arc magmas. For

example, primitive magmas from warm slab subduction zones, such as the Cascades,

Mexico, and western Aleutians, display elevated LREE/HREE (e.g., La/Yb; Walowski et

al, 2015) and coupled high 176

Hf/177

Hf and 143

Nd/144

Nd with lower values of Lu/Hf (Cai

et al., 2014) when compared with arcs associated with older oceanic crust. These

relationships demonstrate that partial melts of subducted oceanic crust play an

increasingly important role in the formation of magmas in arcs beneath which young

oceanic crust subducts.

3.5.6. Model for Melt Generation beneath the Southern Cascade Arc

Evidence provided by the geochemical results outlined above suggests that

southern Cascade arc magmas are produced by a multi-step process involving fluxing of

hydrous slab melts into the hot mantle wedge. Figure 3.9 shows a schematic

interpretation of this process that is based on the thermo-petrologic model results of

Walowski et al., (2015), the shear wave velocity model from Liu et al., (2012) and the

magnetotelluric resistivity data from Wannamaker et al. (2014). In our model, H2O is

retained in the hydrated upper mantle portion of the downgoing slab to greater depths

than that at which H2O is lost from the slab top (Fig. 3.9). Final chlorite breakdown

occurs in the slab interior when the slab top reaches ~75-80 km (Fig, 3.9; Walowski et

al., 2015). At this depth, the upper portions of the slab consisting of MORB-volcanics are

above the MORB+H2O solidus, and thus should melt when fluxed by rising chlorite-

Page 66: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

50

derived fluids (i.e., Skora and Blundy, 2010; Till et al., 2012; Spandler and Pirard, 2013).

The resulting hydrous dacitic melts (Klimm et al., 2008; Jego and Dasgupta, 2013) then

rise into the overlying mantle wedge, and react with the surrounding mantle to produce

hydrous, calc-alkaline, basaltic to basaltic andesite melts.

As a further test of this model of magma formation beneath the southern

Cascades, we determined whether breakdown of chlorite can supply enough H2O to

balance the flux of H2O erupted from the Cascade arc. Previous work in the Nicaraguan

arc and northern Japan has provided evidence for hydration of the deep slab (Ranero et

al., 2003; Garth and Rietbrock, 2014) and the importance of fluids released from the deep

slab in the production of arc magmas (e.g., Spandler and Pirard, 2013). For the Juan de

Fuca plate, higher temperatures at Moho depths caused by the younger slab age may limit

the extent of serpentinization (Nedimovic et al., 2009), but no data are available for the

Gorda plate to assess upper mantle hydration. Because of this uncertainty, Walowski et

al. (2015) conservatively assumed 2 km of hydration below the Moho of the downgoing

plate and a bulk H2O concentration of 2 wt% for the hydrated peridotite. Using these

model parameters, the H2O flux contributed by chlorite breakdown in the slab interior is

estimated to be ~1-2 x 106 kg/km/yr. For the magmatic flux of H2O from the Cascades,

we use the estimate from Ruscitto et al. (2012). This method, which takes extrusive and

intrusive magmatic fluxes into account (Sherrod and Smith, 1990; White et al., 2006) and

utilizes primary volatile contents from the central Oregon Cascades (which overlap with

those from the Lassen region), yields a maximum H2O flux of 1.93 x 106 kg/km/yr. This

agrees very well with the flux resulting from our thermo-petrologic model, demonstrating

that fluids derived from the breakdown of chlorite in the hydrated upper mantle portion of

Page 67: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

51

the slab may be sufficient to produce observed magmatic and volatile fluxes in the

Cascade arc.

The thermo-petrologic model results of Walowski et al., (2015) also predict two

main pulses of fluids from the downgoing slab associated with 1) the final breakdown of

hydrous phases during eclogitization of the oceanic crust, and 2) the final breakdown of

chlorite in the hydrated mantle portion of the slab (Fig. 3.9). The first, more shallow

pulse of fluid release correlates well with the location of low-resistivity anomaly beneath

the forearc (Wannamaker et al., 2014), and likely reflects a region of serpentinization of

the cold nose of the mantle wedge. The second predicted location, caused by the release

of the hydrous melt from the slab, agrees well with regions of low shear wave velocity

beneath the Lassen region (Liu et al., 2012). This is consistent with models of fluid

migration into the mantle wedge which suggest that, for most values of wedge

permeability, slab dip, and convergence velocity, there is a net migration of fluids and

melts away from the trench (Cagnioncle et al., 2007). This implies that arc magmas will

inherit a slab signature from a region of the slab that is slightly up-dip of the region that

lies directly beneath the arc. Therefore, patterns of decreasing amounts of a subduction

component towards the back-arc as observed in the geochemical data (Borg et al., 1997;

2002), provides further support for these interpretations.

Page 68: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

52

Figure 3.9: Schematic interpretation for magma formation beneath the S. Cascades.

Chlorite-derived fluids from the deep slab interior beneath the forearc vents (small blue

arrows) drive flux-melting of the oceanic crust (red colored area), producing hydrous slab

melts that migrate into the overlying mantle (red arrows), where they react with peridotite

to induce further melting. The location of hydrous phase stability in the downgoing slab

(dark blue shaded region) and main pulses of fluid release from the slab (small light blue

arrows) are based on the thermo-petrologic model results of Walowski et al. (2015). Area

of low-velocity (dark and light orange shaded regions for latitudes 41 and 40.6,

respectively) based on shear wave velocity model of Lui et al. (2012). Green shaded

region shows the location of low resistivity from Wannamaker et al. (2014).

3.6. Conclusions

Our results provide strong evidence that magma production in the southern

Cascade arc is driven by hydrous slab melt addition to the mantle wedge. Low B

concentrations and MORB-like B isotope compositions indicate that extensive

dehydration of the plate occurs before it reaches sub-arc depths. However, the abundance

of volatile elements and correlation of volatile and trace element ratios (H2O/Ce, Cl/Nb,

Sr/Nd) show that Lassen region magmas have been enriched by variable amounts of

addition of a hydrous subduction component. Correlation of fluid mobile trace elements

Page 69: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

53

and radiogenic isotopes demonstrates that the subduction component in the southern

Cascades is less radiogenic than the sub-arc mantle wedge, and must be derived from a

partial melt of the subducting Gorda MORB. pMELTS model results show that hydrous-

melt-fluxed melting of the mantle wedge can produce basaltic magmas with similar major

element compositions to those measured in Lassen melt inclusions. Agreement between

pMELTS model results and measured concentrations of H2O and Y confirms that

variations between magmas are likely formed by differences in the amounts of slab

component addition to the mantle wedge at peak temperatures of approximately ~1300-

1350C. Our results provide further evidence that chlorite-derived fluids from the deep

slab interior can flux-melt the oceanic crust, producing hydrous slab melts that migrate

into the overlying mantle, where they react with peridotite to induce further melting.

3.7. Bridge

In the preceding chapter (III), the compositions of olivine-hosted melt inclusions

from eight cinder cones were used provide evidence for slab melting and investigate

magma generation processes in the mantle wedge beneath the southern Cascades. In the

next chapter (III), I will focus on the youngest cinder cone in the Lassen region, Cinder

Cone (1666 C.E., Sheppard et al., 2009) to explore the melt evolution and plumbing

system dynamics prior to and during monogenetic eruptions in the southern Cascades.

Page 70: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

54

CHAPTER IV

UNDERSTANDING MELT EVOLUTION AND THE MAGMATIC

PLUMBING SYSTEM AT CINDER CONE, LASSEN VOLCANIC

NATIONAL PARK, CALIFORNIA: INSIGHTS FROM OLIVINE-

HOSTED MELT INCLUSIONS

This chapter includes material co-authored with Dr. P.J. Wallace, Dr. K. V.

Cashman, J.K. Marks, Dr. M.A. Clynne, and Dr. Philipp Ruprecht. Sample collection was

performed by me with the help of Dr. Wallace, Dr. Cashman, J.K. Marks, and Dr.

Clynne. All melt inclusions were prepared and analyzed by me using Fourier transform

infrared spectroscopy (H2O and CO2), electron microprobe (major elements), and laser

ablation inductively coupled plasma mass Spectrometry (trace elements). Analyses, data

corrections, geochemical modeling, and interpretation were performed by me, with the

help of Dr. Wallace. This melt inclusion dataset is integrated with tephra stratigraphy,

componentry, and textural analyses performed by J.K. Marks (Marks, 2012). I drafted all

the figures used in this article and the text was written by me, with editorial assistance by

Dr. Wallace.

4.1. Introduction

Spanning approximately 1,250 km along the margin of the northwestern United

States and southern Canada, the Cascade magmatic arc is comprised of over 2,300

Page 71: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

55

Quaternary volcanoes (Hildreth, 2007). Less than 30 of these volcanoes, however,

represent the most identifiable high peaks of the Cascades, while the majority of the vents

are small mafic shields and cones (Hildreth, 2007). In the Lassen area alone, over 500

monogenetic cinder cones and small shield volcanoes have erupted in the last 12 Ma

(Guffanti et al., 1990; Borg et al., 1997), and they collectively account for a large fraction

of the total erupted volume in the southern Cascades (Sherrod and Smith, 1990). Despite

the abundance of monogenetic volcanoes in the Cascades, and consequently, their

likelihood to pose volcanic hazards in the northwestern United States, these types of

volcanoes are understudied in comparison to the more evolved and longer-lived

stratovolcanoes.

Compositionally, monogenetic volcanoes in the Cascades are typically mafic, and

they can therefore be used as windows into mantle processes such as arc magma

formation (e.g., Borg et al., 1995, 1997; Leeman et al., 2004; Ruscitto et al., 2010;

Walowski et al., 2015). The eruptions of cinder cones typically last <1-15 years, and

these short timescales often result in limited evolution of their magmas. It has been

shown, however, that despite these short timescales, some cinder cones produce evolved

compositions and complex plumbing systems (e.g., Pioli et al., 2008). At Paricutin in

Mexico, for example, magmas evolved from basalt to andesite over the course of the

eruption that lasted ~9 years (McBirney et al., 1987; Luhr and Carmichael, 1985; Luhr,

2001; Rubin et al., 2004; Erlund et al., 2009). This compositional evolution was

accompanied by a decrease in explosivity and magmatic volatile contents, providing

insight into shallow storage reservoirs beneath cones in central Mexico (Pioli et al., 2008;

Johnson et al., 2008). At Cinder Cone, in Lassen Volcanic NP, CA, bulk compositions

Page 72: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

56

also evolve systematically over the course of the eruption (Clynne et al., 2000). However,

this eruption occurred over a significantly shorter period of time, suggested to be less

than one year (Clynne et al., 2000; Sheppard et al., 2009). How do the compositions of

magmas from Cinder Cone evolve so rapidly? With a complicated inter-fingering of

tephra and lava flows, can we connect pre-eruptive changes in magma chemistry to

changes in eruption dynamics? What are the timescales associated with the formation,

storage, and ascent of these magmas?

To investigate these questions, we focus on the composition of olivine-hosted

melt inclusions sampled from the entire eruptive sequence at Cinder Cone. Melt

inclusions are small volumes of melt trapped inside of phenocrysts at depth, and as such

they can provide two unique insights in addition to traditional methods of bulk rock and

mineral chemistry. (1) Melt inclusions trapped in early crystallizing phases, such as

olivine, provide snapshots of melt chemistry, potentially before additional magmatic

evolution by fractional crystallization, mixing, and/or contamination occurs. In this way,

melt inclusions may provide better information about parental melt compositions, and

thus, better constraints on the processes of formation, storage, and evolution of the

erupted material. (2) Melt inclusions can be used to estimate pre-eruptive volatile

contents (e.g. and Wallace, 2008). The pre-eruptive concentrations of volatiles, such as

H2O, CO2, S, Cl, in a magma are key to understanding a variety of processes from mantle

melting (Gaetani and Grove, 1998; Grove et al., 2006; Hirschmann et al., 2009) to

magma ascent and eruption (e.g., Roggensack, 1997; Metrich and Wallace, 2008). Pre-

eruptive volatile contents are difficult to determine, however, because they decrease in

solubility as pressure decreases and therefore degas from the magma during ascent,

Page 73: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

57

driving explosive eruptions, and leaving the erupted material with only a fraction of the

volatile contents with which it began (Holloway and Blank, 1994). Melt inclusions (MI)

are one of the best tools currently available to directly measure pre-eruptive volatile

concentrations, because they are trapped at depth, often before significant degassing

occurs. Because the solubility of H2O and CO2 in magmas is sensitive to pressure, their

concentrations within MI can also provide estimates of entrapment pressures (Newman

and Lowenstern, 2002; Papale et al., 2006; Iacono-Marziano et al., 2012), and therefore,

magma storage depths (e.g., Johnson et al, 2008; Pioli et al., 2008). Although recent work

has shown that melt inclusions are not perfect storage containers and can lose H+

(protons) by diffusion through the host and CO2 due to formation of a vapor bubble in the

inclusion, there are methods for distinguishing and correcting for these effects (e.g.,

Bucholz et al., 2013; Lloyd et al, 2013; Wallace et al., 2015).

In this study, we use major, trace, and volatile element concentrations in olivine-

hosted melt inclusions from the tephra deposit at Cinder Cone in conjunction with lava

and bulk tephra compositions and olivine major and trace element chemistry. We

interpret the data with regard to stratigraphic relationships between lava and tephra and

textural analyses of tephra clasts (Marks, 2012). We use these results to shed light on the

processes that led to compositional evolution of the magmas and accompanying changes

in eruptive dynamics at Cinder Cone. We show that the volatile contents of melt

inclusions can provide insight into the minimum depths at which these magmatic

processes occur. We also demonstrate the usefulness of melt inclusions in determining

the composition of the mantle source, in a case when the erupted bulk material is

sufficiently contaminated by crustal processes. Lastly, we estimate the timescales of

Page 74: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

58

olivine residence times in the magmatic system to compare with the estimated timescale

of eruption. Using these results, we formulate a schematic model for the plumbing

systems beneath Cinder Cone.

4.2. Geologic Setting

Cinder Cone is a basaltic andesite scoria cone located in the northeast corner of

Lassen Volcanic National Park (Fig. 4.1). Lassen Volcanic National Park encompasses

430 km2 of volcanically-sculpted landscape surrounded by the Sacramento Valley and

Klamath Mountains to the west, the Sierra Nevada to the south, and the Basin and Range

to the east. The park gets its name from Lassen Peak, the tallest point in the national park,

a dacite dome complex that last erupted in 1915. Tectonically, Cinder Cone is found

within the Lassen Segment where volcanism is related to oblique subduction of the Gorda

micro-plate beneath the North American plate, producing the more abundant calc-

alkaline magmas (Clynne and Muffler, 2010). Tholeiitic magmas are also common in this

region, and are often associated with extensional faulting, the results of westward

expansion of the Basin and Range province impinging on the southern Cascade arc. The

quaternary volcanics in the Lassen region sit above older mafic to intermediate volcanoes

and volcanic products 2-4 km thick (Berge and Stauber, 1987). This volcanic basement is

underlain by plutonic Sierran and metamorphic Klamath terrain basement rocks as

suggested by mapped outcropping units (e.g. Saleeby et al., 1989; Cecil et al., 2012) and

constant modeled seismic velocities across the Sierra Nevada-Cascade Range boundary

(Berge and Stauber, 1987).

Page 75: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

59

The eruptions and volcanic history of Cinder Cone have been an ongoing

controversy over the past century, and its eruption in 1666 A.D. was only recently

confirmed by radiocarbon dates of trees killed by the lava flows (Clynne, 2000).

Dendrochemistry and ring-width analyses of living trees were also utilized to determine

that the entire eruption likely occurred in less than one year (Sheppard et al. 2009). At

approximately 350 years old, it is the youngest cinder cone in the Cascade Arc and, as

such, has remained unvegetated allowing for excellent preservation of the lava flows and

tephra deposit (Clynne and Muffler, 2010). The eruptive material is comprised of a ~200-

meter-tall scoria cone built on top of an earlier cone, a tephra deposit as much as 3 meters

in thickness and ~20 by ~10 kilometers in spatial extent, and five main lava flows

separated into three phases – Old Bench, Painted Dunes, and Fantastic Lava (OB, PD,

and FL, respectively; Clynne and Muffler, 2010)(Fig. 4.1).

4.3. Sample Descriptions and Tephra Stratigraphy

The tephra deposit was first described in detail by Heiken (1978) and originally

separated into three phases, Units 1, 2, and 3, thought to correspond with the three phases

of lava flow emplacement, OB, PD, and FL, respectively. Samples used in this study

were collected from a ~1.2-m-deep pit to the north of Cinder Cone (Fig. 4.2; yellow star

in Fig. 4.1). Ten samples were collected from distinctive golden layers throughout the

section. Field descriptions of each sample are listed in Table 4.1, and a corresponding

field photo is labeled in Figure 4.2. The lowermost sample from the column, LCC-1-9, is

interpreted to be the opening phase of Unit 2, because it is thicker than the maximum

Page 76: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

60

measured thickness of Unit 1 by Heiken (1978). Unit 2 is visually distinct from Unit 3 in

that it has appears to contain a higher percentage of larger golden tephra clasts. The Unit

Figure 4.1. Geologic and isopach map of Cinder Cone showing the distribution and

geology of the lava flows, and the whole-deposit tephra isopachs. The main tephra

column used in this study is shown as the yellow star. The three phases of the lava flows

are Purple = Old Bench (tephra Unit 1), orange = Painted Dunes (tephra Unit 2), and

green = Fantastic Lava (tephra Unit 3).

Page 77: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

61

2 tephra deposit, calculated to have a minimum volume of ~0.6 km3, is also less

voluminous and widespread than Unit 3, which covers ~0.15 km3 (Marks, 2012). LCC-1-

5 is a dark fine ash layer which is interpreted to represent the boundary between Units 2

and 3, and also signifies a change in eruptive dynamics (Marks, 2012). The uppermost

layer, LCC-1-1, is overlain by a thin white layer of white 1915 Lassen Peak pumice.

There are several curiosities about Cinder Cone’s deposits and therefore its

eruption. First, there are ubiquitous quartz xenocrysts in all of the basalt-basaltic andesite

lava flows and tephra layers (Clynne et al., 2000). Second, the bulk composition of

erupted material changed from basalt to basaltic andesite through the PD flows and from

basaltic andesite to basalt through the FL flows (Clynne et al. 2011). The tephra mimic

these changes: tephra Unit 2 is compositionally similar to the PD flows and tephra Unit 3

is compositionally similar to the FL flows (Clynne et al. 2011).

Of the ten samples collected, six were chosen for melt inclusion analysis to best

represent the main eruptive phases. Tephra from LCC-9, LCC-7, and LCC-6 represent

the early erupted tephra, LCC-5, and LCC-4 are interpreted to represent the transition in

eruptive phases (closing phase of Unit 2 and opening phase of Unit 3, respectively), and

LCC-2 represents the late erupted tephra (Fig. 4.2). These samples were chosen to

understand the chemical transitions that accompany the textural changes of the tephra

deposit.

Page 78: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

62

Figure 4.2. Stratigraphic section of sample locality. Field photo of the main tephra

column, with samples labeled and color-coded by tephra units. Colors will be used

throughout study. Descriptions can be found in Table 4.1 (Photo credit: K. Cashman.)

Page 79: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

63

Table 4.1. Tephra Stratigraphy and Sample Descriptions Thickness Unit Description

-- -- A thin layer of 1915 pumice lies at the top of the section

10 cm LCC1-1 Coarse ash, vaguely laminated with abundant lava chips

2.5 cm -- Fine ash layer with abundant roots

2.5 cm -- Coarse ash layer

1 cm LCC1-2 Tan fine ash layer

15 cm LCC1-3 Two coarse ash layers separated by a thin fine ash layer

15 cm LCC1-4 Two coarse ash layers separated by a thin fine ash layer

2.5 cm -- Fine ash layer

17.5 cm -- Alternating cm-size layers of coarse and fine ash

2.5 cm LCC1-5 Fine ash layer that separates Fantastic Lava from Painted Dunes

1 cm LCC1-6 Tan Lapilli layer

5 cm -- Coarse ash layer of dark scoria intermingled with blond tephra

3.75 cm LCC1-7 Tan Lapilli layer

3.75 cm -- Dark ash with dense clasts

20 cm LCC1-8 Coarse tan lapilli layer, inversely graded. Base has abundant

oxidized cinder clasts

5 cm -- Dark coarse ash layer

12.5 cm LCC1-9 Coarse ash layer with mixed tan and black clasts

4.4. Methods

4.4.1. Sample Preparation

The ash sized fraction of tephra was used for the melt inclusion analyses because

it quenches more rapidly than lava flows, bombs, and lapilli sized tephra clasts,

minimizing the potential effects post-entrapment modification after eruption (Lloyd et al.,

2013). Loose olivine crystals, 1 mm to 250 μm in size, were hand-picked from the

washed and seived tephra. The olivine crystals were treated in HBF4 to remove adhering

glass and examined in immersion oils (refractive index 1.675) to locate melt inclusions.

Olivine crystals from the deposit are generally euhedral and all contain chrome-spinel

inclusions. Nearly all melt inclusions also contained a 10 to 30-μm-diameter vapor

bubble. Olivine hosting fully enclosed and glassy melt inclusions were mounted in crystal

bond on glass slides and prepared as doubly polished wafers. Wafers were immersed in

acetone to completely dissolve the crystal bond before Fourier transform infrared

Page 80: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

64

spectroscopy (FTIR) analysis, and subsequently mounted in epoxy for electron probe

micro-analysis (EPMA) and laser ablation inductively coupled plasma mass spectrometry

(LA-ICP-MS) analyses.

4.4.2. Analytical Procedures

4.4.2.1. FTIR

H2O and CO2 concentrations were measured at the University of Oregon using a

Thermo-Nicolet Nexus 670 FTIR spectrometer interfaced with a Continuum IR

microscope. Concentrations were calculated from IR peak absorbances using the Beer–

Lambert law: C=MA/ρdε, where M is the molecular weight of H2O or CO2, A is the

measured absorbance of the band of interest (3550 cm-1

for H2O, and 1515 and 1435 cm-1

for CO2), ρ is the glass density at room temperature, d is the thickness of the MI wafer,

and ε is the molar absorption coefficient. An absorption coefficient of 63 L/mol cm

(Dixon et al., 1995) was used for H2O whereas the absorption coefficients for CO2 were

calculated for each inclusion based on the major element composition (Dixon and Pan,

1995). The background subtraction methods used to determine the carbonate peak heights

at 1515 and 1435 cm-1

are described in Roberge et al, (2005).

4.4.2.2. EPMA

Melt Inclusions and host olivine were analyzed for major elements, S, and Cl on

the Cameca SX-100 electron microprobe at the University of Oregon CAMCOR

MicroAnalytical Facility. MI glass compositions and host olivine were analyzed with a

10 nA, 10 μm diameter beam and 15 kV accelerating voltage for major elements. Time

Page 81: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

65

dependent intensity corrections were made for Na, K, Si, and Al, and these elements were

analyzed first. Subsequently, the beam current was increased to 50 nA for collection of S,

Cl, Ti, and P. Individual inclusion analyses are averages of three point analyses. Olivine

compositions are also averages of three point analyses taken ~100 μm away from

inclusions and crystal edges to ensure analysis of olivine unaffected by later stage

diffusion.

4.4.2.3. LA-ICP-MS

Melt inclusions and host olivine were subsequently analyzed for a suite of trace

elements on the Photon Machines Analyte G2 193 nm ArF ‘‘fast’’Excimer Laser system

at Oregon State University, using 50 µm spot size with a 5 Hz pulse rate. Measured trace

element concentrations were determined by reference to GSE-1G glass as a calibration

standard and using 43Ca as an internal standard (see Loewen and Kent, 2012, for details).

BHVO-2G, BCR-2G, and GSD-1G glasses were also analyzed to monitor accuracy and

precision, and the analyzed values were within 10% of accepted values. A larger set of

olivine phenocrysts from LCC-2, LCC-4, and LCC-9 were analyzed at Lamont-Doherty

Earth Observatory. Data was acquired using a 193 ArF Eximer laser with a PQExcell

ICPMS operated in continuous line scan mode using a circular laser spot (25 µm for data

acquisition and 50 µm for pre-ablation) from the core to the rim of each olivine crystal.

Scan speed was 3 µm/s (30 µm/s during pre-ablation) and laser pulse rate was held at 15

Hz. Calibration curves were obtained using a set of reference glasses (BIR, BHVO,

GOR132-G and GOR128-G) for 15 elements (Li-7, Mg-26, Al-27, P-31, Ca-44, Sc-45,

Ti-47, V-51, Cr-52, Mn-55, Fe-57, Co-59, Ni-60, Cu-65, Zn-66). MPI-Ding Gorgona

Page 82: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

66

glasses were specifically included to obtain good calibrations for element compatible in

olivine (most notably Ni). Mg-26 was used as internal standard. New calibration curves

were obtained prior to each sample. Reference glasses KL2-G and ML3B-G, and San

Carlos olivine (USNM 111312/444), were also analyzed to monitor accuracy and

precision.

4.5. Results

4.5.1. Olivine Compositions

The host olivines have core compositions of Fo88-90.5. A larger population of

analyzed olivine phenocrysts from LCC-9, LCC-4, and LCC-2 overlap with the MI host

olivine compositions (Fig. 4.3a), and exhibit broad, homogenous cores and thin Fe-rich

rims. From this larger population, 10 crystals were identified as xenocrysts because they

have variable zonation patterns and contain cores of Fo78-87 . These xenocrysts are

excluded from Figure 4.3 because they are not representative of the main phenocryst

populations. The olivine hosts for the analyzed melt inclusions have high core forsterite

compositions. Early erupted tephra have slightly more Mg-rich olivine cores (Fo89-90.5)

compared with the later erupted tephra (Fo88-89) (Fig. 4.3). Ealy- and late-erupted olivine

phenocrysts are also distinguished from each other in their trace element compositions

(Fig. 4.3b), with early-erupted olivine containing high Fo cores with a more restricted

range in Ni. These values are consistent with olivine previously analyzed from both lava

and tephra (M. Clynne, unpublished data).

Page 83: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

67

4.5.2. Post-entrapment Crystallization and Fe-loss Corrections

The melt inclusions analyzed in this study were all sampled from rapidly

quenched tephra. However, melt inclusions are commonly affected by crystallization of

olivine along the walls of the inclusion and by Fe diffusive loss during the time between

trapping and eruption. If the measured composition of a melt inclusion is not in

equilibrium with the surrounding host olivine, then the melt inclusion has likely

experienced post-entrapment crystallization (PEC; Danyushevsky et al., 2000). The

original melt inclusion composition can be estimated by incrementally adding

equilibrium olivine back into the melt until it is in equilibrium with its host

Figure 4.3. Olivine phenocryst compositions. a) Fosterite content of olivine hosting

melting inclusions (filled symbols) and larger olivine population measured by LA-ICP-

MS (open symbols) ordered by relative stratigraphic height. See Table 4.1 and Fig. 4.2

for exact stratigraphic heights. b) Ni versus Fo for olivine phenocryst cores from the

larger population measured by LA-ICP-MS. This excludes olivine xenocrysts, as

described in main text. Solid black lines are fractionation curves labeled with tick marks

which represent 1% of olivine fractionation.

Page 84: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

68

olivine (e.g., Ruscitto et al., 2010). In this study, Petrolog 3.1 (Danyuchevsky and

Plechov, 2011) was used to determine original melt inclusion compositions and to

calculate the percent of olivine crystallized from the melt inclusion after entrapment. For

this correction, Petrolog 3.1 used Fe-Mg partition coefficients (Kd) between the melt and

host olivine from Ford et al. (1983). Oxygen fugacity was determined using the

partitioning of V between the melt inclusion and host olivine using methods of Mallmann

and O’Neill (2009). The oxygen fugacities of the most primitive MIs were ΔQFM +1.4

(±0.4), with no distinct difference between eruptive units.

Post-entrapment crystallization corrections also require an estimate of the initial

Fe content of the MI prior to Fe-loss by diffusion. Bulk tephra analyses from Cinder

Cone were used as a proxy to estimate the original Fe contents. Primitive lavas in the

Lassen region have 7-8 wt% FeOT which overlaps with or is slightly higher than the

values for Cinder Cone lavas and bulk tephra (5.5-7 wt% FeOT). Based on this

comparison we assumed that Cinder Cone melts initially had 7 wt% FeOT. Volatile and

incompatible trace element concentrations in the melt inclusions were also corrected for

PEC using the Petrolog 3.1 results.

4.5.3. Major Element Compositions

We present data for 47 melt inclusions, 24 of which are from early units, 15 from

transitional units, and 8 from the late-erupted tephra. Corrected melt inclusions are

basaltic to basaltic andesite in composition (Fig. 4.4), and most fall in a narrow range of

SiO2 contents (51-53 wt% on a volatile-free basis). Variability in total alkalis, however,

shows compositional differences in MIs from different units. Early-erupted units (LCC-9,

Page 85: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

69

LCC-7, LCC-6) have lower alkalis (~3.5-4 wt% Na2O+K2O) compared to middle- and

late-erupted units (LCC-5, LCC-4, LCC-2), which have higher values (~4-4.5 wt%).

Middle-erupted or transitional units (LCC-4 and LCC-5) display the most compositional

variability and contain the greatest number of inclusions that have higher concentrations

of both SiO2 and alkalis. This pattern is also apparent in the concentrations of other major

elements, such as MgO and CaO, which are elevated in early-erupted MIs compared to

late-erupted MIs, but highly variable in the transitional units (Fig. 4.5). Because early-

and late-erupted MI are easily distinguished with most major elements and are found

within tephra units that display distinct eruptive dynamics (Marks, 2012), for simplicity,

we will refer to the average early-erupted MI composition as magma batch 1 and the

average late-erupted MI composition as magma batch 2.

Most MIs have distinctly lower SiO2 than the lavas and bulk tephra (Fig. 4.4).

This is an indication, as will be discussed further below, that the magmas at Cinder Cone

were crustally contaminated after olivine growth and MI entrapment. Furthermore, most

olivine crystals (Fo89-90) are not in Fe-Mg equilibrium with the lava and bulk tephra and

often contain thin reaction rims of pyroxene (M. Clynne, unpublished data). We did not

analyze matrix glass, but the composition of olivine in equilibrium with the bulk tephra

(which is olivine enriched compared to the matrix) is Fo87-87.5. The olivine phenocrysts

rims are also commonly Fo88-87. Therefore, most olivine phenocrysts at Cinder Cone must

have grown from a more primitive parental magma before the magma became

contaminated. Interestingly, the differences in alkalis between early-erupted and late-

erupted MIs are mirrored in the total alkali concentrations of lava and bulk tephra (Fig.

4.4). Melt inclusions with elevated SiO2 (>54 wt%), which mostly come from the

Page 86: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

70

transitional units, overlap with lava and tephra in MgO and Al2O3 content, and likely

represent melts that were affected by crustal contamination before being trapped in

olivine (Fig. 4.4 and 4.5).

4.5.4. Volatile Concentrations

H2O and CO2 concentrations measured in melt inclusions at Cinder Cone range

from 0.7-3.5 wt% and 0-1500 ppm, respectively (Fig. 4.6), similar to values for other

Figure 4.4. Melt inclusion total alkalis and SiO2 compared to bulk tephra and lava. Filled

symbols are melt inclusions from this study normalized on a volatile-free basis. Bulk lava

(open diamonds) and tephra (crosses) are unpublished data from M. Clynne.

Classification boundaries for basalt, basaltic andesite, andesite, trachy-andesite, and

trachy-basalt from LeBas et al., (1986). Uncertainty is less than the symbol size (See

Supplementary Table C.1; see Appendix C for all supplementary information for this

chapter).

Page 87: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

71

Figure 4.5. Melt inclusion MgO vs. Al2O3 and CaO. MgO vs. a) Al2O3 and b) CaO for

MI compositions (normalized on a volatile-free basis) compared with lava and bulk

tephra (shaded blue region; M. Clynne unpublished data). Fractionation curves calculated

using Petrolog 3.1 for olivine (Ol, black dashed line; Danyushevsky and Plechov, 2011)

and Rhyolite-MELTS at 10 kbar for clinopyroxene (CPX, grey line; Gualda et al., 2012),

with tick marks representing 1% fractionation. Uncertainty is less than symbol size (see

Supplementary Table C.1).

Page 88: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

72

nearby cinder cones Walowski et al., 2015; in prep.) and cones in the central Oregon

Cascades (Ruscitto et al. 2010; Ruscitto et al., 2011; LeVoyer et al., 2011). However, a

number of recent studies have shown that MI commonly experience post-entrapment

changes that decrease their dissolved H2O and CO2 concentrations (Esposito et al., 2011;

Gaetani et al., 2012; Bucholz et al., 2013; Lloyd et al., 2013; Hartley et al., 2014; Moore

et al., 2014; Wallace et al., 2014). The measured dissolved concentrations therefore

represent a minimum estimate of the initial H2O and CO2 contents of the magma during

entrapment.

Most basaltic MI, including nearly all MI in this study, contain a vapor bubble,

the size of which can be used to determine if it is primary (co-entrapped with melt) or

formed after entrapment (Riker et al., 2005; Aster et al., in prep.). Shrinkage bubbles

form after entrapment as a result of the pressure drop caused by crystallization along the

host-inclusion interface and thermal contraction of the melt (Roedder, 1979; Lowenstern,

1995). Because CO2 has a much lower solubility than H2O, it strongly partitions into the

bubble, leaving only a fraction of the CO2 dissolved in the melt compared to that at the

time of entrapment (Esposito et al., 2011; Hartley et al., 2014; Moore et al., 2014).

Recent work suggests that 40-90% of the total CO2 in the MI can be lost to this shrinkage

bubble (Moore et al., 2014; Wallace et al., 2014), leading to a drastic underestimation of

entrapment depths when only dissolved CO2 is taken into account. To correct for CO2

loss, we used a modified calculation based on the methodology of Riker (2005) and

Wallace et al. (2015). A comparison of this method with results of experimental

redissolution of CO2 in shrinkage bubbles and Raman spectroscopic determination of

CO2 densities in bubbles shows good agreement (Wallace et al., 2015; Aster et al., in

Page 89: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

73

prep). First, we estimate the size of the vapor bubble at the time of eruption (Riker,

2005). This bubble volume is formed as the result of cooling-induced crystallization

(PEC) and melt contraction as a function of the change in temperature between

entrapment and pre-eruption (∆T). We use the phase equilibria of the most primitive melt

inclusion compositions calculated with rhyolite-MELTS (Gualda et al., 2012) and the

volume and thermal expansion data for silicate melts (Lange and Carmichael, 1990) and

olivine (Kumazawa and Anderson, 1968; Suzuki, 1975) to derive an equation for the

Cinder Cone magmas:

Bubble volume % = 0.0092 ∆T. Equation (1)

For each individual melt inclusion, we used Petrolog 3.1 to estimate the ∆T and then

equation (1) to estimate bubble volume %. Melt inclusions that did not contain a vapor

bubble were excluded from this correction. Although the absolute temperatures

calculated with Rhyolite-MELTS (Gualda et al., 2012) and Petrolog 3.1 (Danyushevsky

and Plechov, 2011) are different, the ∆T calculated for each MI from the two methods is

similar (Aster et al., in prep).

With this estimate of the pre-eruption bubble volume %, the pre-eruption

temperature, and the dissolved H2O and CO2 concentrations of each inclusion, we used

the VolatileCalc program (Newman and Lowenstern, 2002) to determine the pressure of

entrapment and the mol% CO2 in the equilibrium vapor phase. We then used the Redlich-

Kwong equation of state to determine the molar volume of CO2 in the vapor bubble. The

results of the CO2 restoration calculations suggest that 25-78% of the initial dissolved

CO2 in the MIs at the time of trapping was lost to vapor bubbles during post-entrapment

crystallization. Corrected concentrations of CO2 are shown in Figure 4.6.

Page 90: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

74

The H2O contents of MIs can also be modified after entrapment. Experimental

studies have shown that at magmatic temperatures, olivine hosted MI can lose nearly all

of their H2O in hours to days via H+ (proton) diffusion, depending on temperature

(Gaetani et al., 2012; Bucholtz et al., 2013). To investigate whether some MI from Cinder

Cone had lost H by diffusion, we compared S/K2O and H2O/K2O ratios, because

correlated low values would indicate that lower H2O values in the MI were the result of

degassing of the melt before it was trapped (e.g., Johnson et al., 2010). However, the

inclusions show variable H2O/K2O with a more restricted range of S/K2O, and this

suggests the H2O variations may be the result of variable diffusive loss. For this reason,

we cannot employ the methods described in Lloyd et al. (2013) to infer original H2O

contents. Thus, to correct for post-entrapment H loss by diffusion, we assumed that all

melt inclusions initially had H2O/K2O values similar to that of the highest measured

value. This correction procedure should yield the maximum likely initial H2O content for

each inclusion. Melt inclusions that have >1 wt% K2O were not corrected for H-loss by

diffusion because they likely experienced crustal contamination, and may have been

trapped more shallowly after some degassing of H2O.

Trapping pressures for the restored H2O and CO2 concentrations were calculated

using solubility relations of Iacono-Marziano et al. (2012) and Papale et al. (2006). The

two methods gave similar trapping pressures, with values ranging from ~2-4.2 kbar for

the corrected MI (Fig. 4.6). These estimates are nearly 2 kbar higher than those estimated

from measured dissolved H2O and CO2 contents. Assuming an average crustal density of

2.6 g/cm2, these pressures represent entrapment depths of ~7-15 km below the surface.

Three melt inclusions did not contain vapor bubbles and were trapped more shallowly

Page 91: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

75

(<2 kbar). Corrected melt inclusions fall more closely along open- or closed-system

degassing paths (Fig. 4.6), indicating that they may have been trapped at a range in

pressures, but this could simply be an artifact of our assumption of initially constant

H2O/K2O values (cf., Lloyd et al., 2012).

Figure 4.6. Melt inclusion corrected H2O vs. CO2. H2O wt% and CO2 ppm for corrected

(filled symbols) and measured/uncorrected (open symbols) MI, as described in the text

(Supplementary Tables C.1 and C.2). Inclusions without vapor bubbles do not require

this correction and were trapped at lower pressure. Isobars calculated using methods of

Papale et al., (2006) for an average melt inclusion composition. Open and Closed system

degassing paths calculated using the VolatileCalc program (Newman and Lowenstern,

2002). Error bars are one standard deviation.

4.5.5. Trace Element Concentrations

Melt inclusions at Cinder Cone have typical calc-alkaline, subduction-related

trace element patterns that are enriched in large ion lithophile (LILE) and light rare earth

(LREE) elements relative to high field strength elements (HFSE) like Nb and Ta (Fig.

4.7). MI from late erupted units have higher concentrations of highly incompatible

Page 92: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

76

elements than early erupted MI (Fig. 4.7). MIs from transitional units that have elevated

SiO2 and alkalis are even more enriched in highly incompatible elements such as Sr, Ba,

U, and Pb (Fig. 4.7).

Figure 4.7. Spider diagram of average melt inclusions compositions. MI trace element

compositions corrected for PEC and normalized to depleted MORB mantle (DMM;

Workman and Hart, 2005). Blue and red shaded regions show the range in trace element

compositions for late- and early-erupted tephra, respectively. Black lines represent

individual melt inclusions from LCC-4 and LCC-5 that have higher SiO2 contents to

highlight trace element enrichments seen in these units.

4.6. Discussion

4.6.1. Crustal Contamination at Cinder Cone

The major element compositions of most MIs have distinctly lower SiO2 than the

lava and bulk tephra at Cinder Cone. This observation suggests significant chemical

evolution of the magma after olivine growth and melt inclusion entrapment. The

ubiquitous presence of quartz xenocrysts in the tephra and lava suggests that this

evolution was dominantly the result of crustal contamination. Partially melted and/or

Page 93: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

77

vesiculated granitic clasts in the lavas (M. Clynne, unpublished data) likely represent the

source of the abundant quartz xenocrysts and rare K-feldspar xenocrysts.

Figure 4.8 shows the results of mixing calculations between the low-SiO2 melt

inclusion compositions and high-SiO2 compositions of the granitic xenoliths. Bulk tephra

and lava compositions fall along mixing lines between magma batch 1 (early-erupted MI)

and magma batch 2 (late-erupted MI) and the average granitic xenolith composition.

Early erupted PD lava and tephra can be produced by 10-35 wt% assimilation of granitic

basement by magma batch 1 (Fig. 4.8). Late erupted FL lava and tephra can be produced

by 15-30% assimilation of granitic basement by magma batch 2 (Fig. 4.8). Rare mafic

enclaves in the late-stage FL lava flows are similar in composition to the LCC-2 MI

compositions (M. Clynne, unpublished data), providing further evidence that the MI

represent the parental magma compositions. Unit 1 tephra, associated with the eruption of

the Old Bench lava flows, is compositionally similar to mildly contaminated early-

erupted Unit 2 tephra (which contains uncontaminated MI – samples LCC-9, LCC-7, and

LCC-6).

Melt inclusions from LCC-5 have TiO2 values that are more similar to the early-

erupted units or that fall along mixing lines with the early-erupted units. MI from LCC-4,

conversely, have TiO2 contents more similar to the late-erupted tephra. These

associations are consistent with our interpretations of the tephra stratigraphy, which

suggest that LCC-5 represents the final, waning phase of Unit 2, whereas LCC-4 is the

opening phase of Unit 3 (Fig. 4.2). LCC-5 also has the largest number of MIs that have

contaminated compositions, and which are similar to the lava flows and bulk tephra. This

may be related to the slowing of the eruption, which lead to shallow crystallization as the

Page 94: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

78

Figure 4.8. Modeled mixing with granitic xenoliths. SiO2 vs. a) TiO2 and b) MgO

comparing MI compositions (normalized on a volatile-free basis; filled symbols) and

lava, bulk tephra, and quenched mafic enclaves (M. Clynne unpublished data). Curves

represent bulk mixing between granitic xenolith compositions (M. Clynne unpublished

data) and average composition of batch 1 (early-erupted MI; red dashed curve) and batch

2 (late-erupted tephra; blue dashed curve) parental magmas, see text for details. Tick

marks represent increments of 10% bulk addition of granitic material. Average analytical

uncertainty in upper right hand corner of each panel (see Supplementary Table C.1).

Page 95: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

79

first eruptive phase came to a close. This slowing or waning of the first eruptive phase is

evidenced by increased groundmass crystallinity seen in the transitional tephra clasts

(Marks, 2012).

Further evidence for assimilation of granitic basement is provided by the trace

element compositions of MIs. Figure 4.7 shows that MI with elevated SiO2, which are

dominantly found in the transitional units, also have higher concentrations of

incompatible trace elements, such as Pb, Ba, U. These incompatible trace elements are

also found in high concentrations in the xenoliths and locally outcropping granitic

basement rocks (Wolf Creek Granite; M. Clynne, unpublished data), the likely source of

the xenoliths. The Li concentrations of the MIs appear to be particularly sensitive to the

granitic contaminant, correlating linearly with Pb (Fig. 4.9), and clearly separating MI

that have experienced contamination from those that did not.

4.6.2. Parental Melt Compositions

The olivine phenocrysts in the Cinder Cone tephra have compositions of Fo88-90.8

and the host crystals for the MI range up to Fo90.2. The upper end of the range of olivine

compositions overlaps with olivine that would be in equilibrium with mantle peridotite.

The trapping pressures of the MI, after accounting for CO2 lost to vapor bubbles and

possible H+ losses, are in the range of 2 to 4.2 kbar, providing evidence that the host

olivine crystallized in the middle crust. These observations require that either parental

melts in equilibrium with the mantle ascended into the middle crust without fractionating,

Page 96: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

80

or that mantle sources beneath the Lassen region are relatively refractory, with olivines

>Fo90.

Figure 4.9. Melt inclusion Li vs. Pb corrected for PEC. High Li and Pb concentrations in

MI are indicative of addition of granitic material, and highlight melt inclusions which

sampled the contaminated magma. Errors can be found in Supplementary Table C.1.

Same symbols as in Fig. 4.8.

Late-erupted MI from LCC-2 show distinct compositional differences when

compared to early-erupted MI (Figs. 4.4 and 4.6), and are hosted in slightly less-

magnesian olivine (Fo88; Fig. 4.3), indicating that they could be related to the early-

erupted magmas via crystal fractionation. We used Petrolog 3.1 and pMELTS (Ghiorso et

al., 2002) to model equilibrium crystallization of the early-erupted magma at 4 kbar, the

maximum trapping pressure estimated from the corrected CO2 and H2O contents (Fig.

4.6). Olivine is predicted to be the liquidus phase for the primitive batch 1 composition,

but olivine crystallization alone cannot account for the decrease in CaO with decreasing

MgO that is observed (Fig. 4.5). Because these basaltic magmas are ultimately derived

from the mantle, they may have stalled near the base of the crust and undergone an

episode of higher pressure crystallization before ascent into the crust. Pressures near the

Page 97: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

81

Moho beneath the southern Cascades are ~10 kbar (~38 km; Mooney and Weaver, 1989)

a pressure above which clinopyroxene may replace olivine as the liquidus phase in

hydrous primitive basalts (e.g., Weaver et al., 2011; Blatter et al., 2013). Mass balance

modeling suggests that ~5% high-pressure clinopyroxene fractionation may explain the

observed major element differences between the two units (Fig. 4.5). However, phase

equilibrium calculations with pMELTS at 10 kbar for the composition of the early-

erupted magmas do not predict clinopyroxene as the liquidus phase, so we test this

possibility further using the trace element compositions of the uncontaminated melt

inclusions. Using equilibrium crystallization and partition coefficients between

clinopyroxene and basaltic melt for La, Nb, Sc and Th, we model how the melt

composition will change as the result of clinopyroxene fractionation. During

clinopyroxene fractionation, Sc is compatible and should decrease rapidly in the melt,

whereas Nb, Th, and La are incompatible, and should increase. For small amounts of

olivine and/or clinopyroxene fractionation, Nb, Th, and La will only increase slightly and

show little variability (Fig. 4.10 b, c, d). Figure 4.10 shows that 5% clinopyroxene can

explain the variability in Sc and the relatively constant Th concentrations between early-

and late-erupted units. However, the magnitude of Nb and La increases with decreasing

MgO wt% cannot be explained by fractionation alone. Thus, we suggest small amounts

of high-pressure clinopyroxene fractionation have likely affected the compositions of

batch 2 parental magmas, but conclude this processes cannot account for the differences

in trace element abundances between them. Because magma batches 1 and 2 are not

related solely by fractional crystallization, either in the mantle or in the crust, their subtle

Page 98: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

82

differences are likely the result of mantle processes, such as variability in mantle source

composition or the degree of melting.

Figure 4.10. Melt inclusion MgO vs. trace elements. MgO (normalized on a volatile-free

basis) vs. a) Sc, b) Th c) La and d) Nb (PEC corrected) compared with lava and bulk

tephra (blue shaded region; M. Clynne unpublished data). Panels c) and d) exclude MI

with elevated Li and Pb, and determined to be contaminated. Fractionation curves

calculated using Petrolog 3.1. for MgO and the equilibrium crystallization equation and

appropriate partition coefficients for trace elements for olivine (Ol, solid grey line) and

clinopyroxene (CPX, dashed black line) with tick marks which represent 1%

fractionation. Partition coefficients were taken from Paster et al., (1974; Sc, CPX-basalt),

Beattie et al., 1994 (Sc, basalt-Ol, La, basalt-Ol), Hauri et al., (1994; Th., basalt-CPX, La,

basalt-CPX, and Nb, basalt-CPX), and McKenzie and O’nions, (1991; Th, basalt-OL, and

Nb, basalt-Ol). Average analytical uncertainty designated by symbol it top right corner of

each panel. Same symbols as in Fig. 4.8.

Page 99: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

83

Compositional variability amongst olivine phenocrysts can further be used to

distinguish between early- and late-erupted parental magmas. Figure 4.3 shows the

olivine Ni and Fo core compositions and reveals differences in early- and late-erupted

units. LCC-9 has olivine phenocrysts cores with high Fo and a more restricted range in

Ni when compared with the late erupted units. Late-erupted olivine cores follow a steep

fractionation path (~2-3 % olivine fractionation) beginning with higher Ni, requiring

either a more refractory mantle source or presence of mantle pyroxenite (Straub et al.,

2011). The range in Ni of early-erupted olivine cores requires less than 1% fractionation

of olivine. Interestingly, some olivine phenocrysts from LCC-9, the earliest eruptive

phase, overlap with olivine from LCC-2 and LCC-4 (Fig. 4.3b). Given this similarity, it

seems likely that the lower-Fo olivine in LCC-9 were derived by interaction of batch 1

and 2 magmas (the latter represented by LCC-2 or LCC-4).

It is important to note that although magma batches 1 and 2 are distinct, they are

more similar to each other than to most other primitive basaltic magmas in the Lassen

region (Fig. 11; Clynne, 1995; Borg, 1995; Walowski et al., 2015). Walowski et al. (in

prep), show that most primitive magmas in the Lassen region are formed by variable

amounts of slab component addition to a relatively heterogeneous mantle wedge. Figure

4.11 shows that early- and late-erupted magmas have similar values of Sr/Nd, indicative

of similar amounts of subduction component addition (Walowski et al., in prep).

Differences in Nb/Zr, however, suggest that late-erupted magma batch 2 may be derived

from a slightly more enriched mantle source (Fig. 4.11). It is also possible that both

batches were derived from the same mantle source, but separated and experienced

different paths, down-temperature, through the upper-most lithospheric mantle. For

Page 100: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

84

example, magma batch 2 may have reacted with pyroxenite veins in the uppermost

lithospheric mantle, which could explain the elevated Ni contents in olivine compared to

magma batch 1. Together, the observations demonstrate that monogenetic cones may tap

different batches of mantle-derived melts during a single, short-lived eruption (Brenna et

al., 2010, 2011; McGee et al., 2011). How these different magmas were tapped and where

they were stored, however, requires further analysis.

Figure 4.11. Comparison of Cinder Cone to Lassen region magmas. PEC corrected MI

compositions of Sr/Nd and Nb/Zr compared with other primitive lavas in the Lassen

Region (Borg et al., 1997). Symbol size for MI exceeds analytical uncertainty. Same

symbols as in Fig. 4.8.

4.6.3. Timescales of Mixing and Ascent

Olivine phenocrysts from all units typically have broad, homogeneous cores, and

Fe-enriched rims (Fig. 4.12). Because Ni is compatible in olivine and has a similar

diffusivity (Petry et al., 2004), it displays a similar pattern to Fe/Mg in the Cinder Cone

Page 101: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

85

olivine phenocrysts. These Fe-rich and Ni-poor rims were likely developed during crustal

contamination, magma mixing or entrainment of olivine in new batches of melt,

processes that could lower the Mg and Ni content of the bulk magma. Although some

new olivine growth may have occurred, we can use the diffusion timescales of trace

elements, such as Ni, to estimate a maximum timescale of olivine residence in an evolved

magma prior to eruption.

The effect of diffusion on the Ni concentration in an olivine crystal can be

calculated with Fick’s second law, which describes the change in composition with time

over a one-dimensional system:

, Equation (2)

where C is concentration, t is time, and D is the diffusivity. The above expression of

Fick’s law assumes a constant diffusion coefficient, D, which for most natural systems is

an approximation. The diffusion coefficient for Ni in natural olivine, for example, varies

with temperature, pressure, oxygen fugacity and composition (Petry et al., 2004; Dohmen

& Chakraborty, 2007). We also assume that little to no olivine growth occurred during

the mixing event and as a result, our estimate will be a maximum timescale because

olivine growth is faster than Ni diffusion. The diffusion coefficient for Ni in the olivine

phenocrysts is assumed to be 2.6 x 10-16

m2/s, taken from Petry et al. (2004) for a

temperature of 1200C. For our boundary condition, we assume local equilibrium, that is,

the measured Ni content at the olivine rim is the equilibrium Ni content of the olivine

coexisting with the melt surrounding each olivine crystal. We also assume that the

average core composition is representative of the initial olivine core composition before

magmatic evolution (Fig. 4.12). To determine the best-fit diffusion curve, we calculated

Page 102: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

86

diffusion profiles for a series of time-steps. The average of the two timesteps that were

visually determined to best bracket the measured Ni concentration gradient at the olivine

rim was used as the best timescale estimate (Fig. 4.12).

Figure 4.12. Ni diffusion in olivine. Panels a) and b) show examples of diffusion model

results as described in the text. Data points represent measured core-rim Ni contents

measured by LA-ICP-MS (symbols are the same as in Fig. 4.8). Model curves for 20

(black dashed curve) and 40 (solid black curve) days typically best fit the data. Panel c)

shows a probability density function of average model timescales for 7 different olivine

crystals from LCC-9, 6 from LCC-4, and 6 from LCC-2.

With these methods, we find that olivine phenocrysts from all units have an

average residence time of 20-40 days. Some phenocrysts have longer apparent

timescales, up to ~90 days, but uncertainty related to our model assumptions may cause

this discrepancy for a few olivine crystals. The largest source of uncertainty in this

method is derived from choice in temperature, and thus, diffusivity. Petrolog 3.1

Page 103: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

87

estimates average trapping temperatures of ~1225C and pre-eruptive temperatures of

~1150C. This indicates that temperatures were decreasing during the mixing event,

which is not captured by our model. Decreasing the temperature would result in longer

timescales, however, our model also does not include olivine growth, which would

decrease the timescale.

The calculated residence times at Cinder Cone are much more rapid than those

calculated for olivine at Jorullo volcano in central Mexico (Johnson et al., 2008), which

suggested olivine residence times of 200 to more than 1000 days. This is consistent with

the total length of the eruption, as the eruption of Jorullo lasted for approximately 15

years (Luhr and Carmichael, 1985), whereas the Cinder Cone eruption probably lasted

less than 1 year (Clynne et al., 2000, Sheppard et al., 2009). Differences between mixing

and storage timescales and the eruption durations at Cinder Cone and Jorullo emphasizes

the diversity of complex behaviors that can be displayed at monogenetic cones. Although

there are large uncertainties associated with the calculated timescales, they still provide

some constraints for the processes associated with magmatic evolution prior to the

eruption of Cinder Cone. Importantly, our results demonstrate that the parental magmas

mixed with granitic crustal material very rapidly prior to eruption.

4.6.4. Mechanisms of Crustal Contamination and Plumbing System Evolution

Although the role of crustal contamination in formation of the erupted magmas at

Cinder Cone is clear, the mechanism by which this contamination occurred is more

difficult to decipher. Determining how contamination occurred can provide constraints on

how the magmatic plumbing system evolved prior to and during eruption.

Page 104: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

88

Several, somewhat contradictory observations suggest two main hypotheses that

may explain the contamination of the Cinder Cone magmas: 1) A shallow, strongly

contaminated batch of magma, perhaps andesitic, was mobilized by and mixed with new

batches of magma coming from depth, or 2) an initial batch of magma thermally pre-

conditioned the granite, creating a zone of crystal- and xenolith-rich rhyolitic mush that

was quickly incorporated into the rising new primitive magma batches. In hypothesis (2),

there is no intermediate magma.

An important observation in support of hypothesis (1) is that granitic

contamination is ubiquitous, but the amount of contamination is heterogeneous and

changes throughout the course of the eruption. Specifically, Figure 4.13 shows the

temporal evolution of the SiO2 contents of the erupted tephra. For both eruptive phases,

magmas start off more evolved (andesitic), and become progressively less contaminated

over time. These temporal patterns in magma composition are similar to those observed

during eruptive episodes 2-47 of Pu’u ’O’o in Hawaii (Garcia et al., 1992; Wallace and

Anderson, 1998), and may indicate that a shallow, evolved magma batch was

simultaneously mixed with and pushed out of the conduit system by the ascending mafic

magma.

However, mineralogical or textural evidence for this contaminated magma seems

to be lacking. Olivine xenocrysts with lower-Fo cores may be representative of the

shallowly stored evolved magma. However, some of these xenocrystic olivine display

low-Fo cores and high-Fo shoulders (Costa and Dungan, 2005), as well as low-Ni cores,

high-Ni shoulders, and normally zoned rims, suggesting that they began to equilibrate

with the primitive parental magma before crustal contamination occurred (Fig. 4.14).

Page 105: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

89

Two xenocrysts from unit LCC-4 do not have high-Fo rims, and have

significantly lower Fo contents than the other xenocrysts, and may be representative of

the mixed magma (Fig. 4.14). Conversely, these same crystals may have grown in the

contaminated magma during the waning and transitional period of the eruption, as they

are only observed in the transitional unit, LCC-4. In addition, we did not analyze olivine

less than 250 µm in diameter, and there could be a population of low-Fo olivine derived

from the contaminated magma in this size fraction. None of the analyzed MI have low-Fo

olivine hosts.

Figure 4.13. Temporal variability in bulk tephra compositions. All data in this figure is

unpublished data from M. Clynne, and depicts how tephra compositions change over the

course of the eruption at Cinder Cone. Unit transitions as defined by Heiken, (1978). The

Unit 1is equivalent to Old Bench. The opening phase of Unit 2 is equivalent to LCC-9.

The Unit 2-Unit 3 transition is equivalent to the transition between LCC-4 and LCC-5.

Melt Inclusions were not samples from Unit 1.

Page 106: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

90

Figure 4.14. Core to rim variations in olivine xenocryts from LCC-9 (early-erupted; red),

LCC-4 (transitional; green), and LCC-2 (late-erupted; blue) tephra.

In support of hypothesis (2), lava and bulk tephra fall on mixing trajectories

between the granitic material and the two parental magma batches. If, for example,

magma batch 2 evolved as the result of mixing with the first batch of contaminated

erupted magma, one would expect mixing trajectories between magma batch 2 and the

contaminated magma of early-erupted units, which is not observed.

Using the combined observations presented above, we suggest the following

model for plumbing system evolution at Cinder Cone. First, one or several intrusions of

batch 2 magma enter the middle crust and form a set of small sills, which may be

partially integrated, over a range of depths. These batch 2 intrusions crystallize olivine,

and have variable compositions and trapping depths as evidenced by fractionation trends

in Fo vs. Ni (Fig. 4.3b). In a manner analogous to the model proposed by Grove et al.,

(1998) at Burnt Lava Flow, which involves the separation of heat and mass transfer, the

mid-crustal sills quench on their margins, and become mechanically separated from the

Page 107: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

91

surrounding granitic basement. However, the sills may transfer heat to the surrounding

granitic wall rocks, causing them to variably melt or disaggregate. The shallowest sill or

sills then become contaminated, where some MI with contaminated compositions and

overlapping to lower trapping pressures are derived. Some of the unzoned xenocrysts

may also be sourced from various parts of this sill complex, such as the Fo77-79 (Fig.

4.14), that produced highly contaminated magma.

Batch 1 magma then rises from the deep crust or mantle. It interacts at least

partially with the existing sills because it samples olivine phenocrysts from batch 2

magmas, as suggested by Fo vs. Ni (Fig. 4.3b). A small portion of batch 1 magma

interacts briefly with the partially melted granite or contaminated magma, and erupts as

Old Bench lava flow and Unit 1 ash (these magmas are compositionally similar to early-

erupted lava and tephra; Fig. 4.8). Most of batch 1 magma stalls when it interacts with the

partially melted granite, mixing and crystallizing rims for 20-40 days. At this point, the

entire system becomes destabilized, and eventually erupts, with the most heavily

contaminated parts erupting first. As a result, the batch 2 sills also become destabilized,

causing them to ascend, interact with heated and partially molten granite, crystallize

lower Fo rims for !20-40 days, and finally erupt in a similar manner to the previously

erupted batch 1 magmas.

4.7. Conclusions

The eruption of Cinder Cone produced a series of tephra deposits and lava flows

that display complex changes in chemistry over the course of the eruption. High-Fo

olivine phenocrysts from all erupted units contain melt inclusions that are more primitive

Page 108: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

92

in composition than the erupted material. The evolved compositions of the lava and bulk

tephra and the abundance of quartz xenocrysts within the deposits suggest the parental

magmas were rapidly contaminated by granitic material in the middle to upper crust. The

basaltic parental magmas sampled by the melt inclusions crystallized at minimum depths

of 2-4.2 kbar, which equates to approximately 7-15 km below the surface. Distinct

compositional variability between early- and late-erupted units suggest two different

mantle-derived basaltic magmas were tapped during the eruption. These two basaltic

compositions correlate with the two main explosive phases of the eruption, and were

separated by an eruptive slowing or pause evidenced by textural analyses of tephra clasts

(Marks, 2012). Diffusion modelling of Ni gradients in olivine rims suggests that olivine

residence times in an evolving magma were on the order of weeks, shorter than those

calculated for longer-lived cinder cone eruptions, such as Jorullo in Mexico. Temporal

changes in lava and bulk tephra compositions indicate that the granitic material was

heterogeneously incorporated after olivine growth and melt inclusion entrapment.

Although the geochemical evidence suggests that the parental magmas were

contaminated by either a shallowly stored evolved magma or a granitic mush, the exact

mechanisms that drove crustal contamination of the Cinder Cone magmas remain

enigmatic. Our combined results provide new insight into the complexities of short-lived

monogenetic eruptions.

Page 109: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

93

APPENDIX A

CHAPTER II SUPPLEMENTARY MATERIALS

A.1. Methods

Cinder cones were chosen based on their primitive compositions to ensure

minimal contributions from crustal processes. At each cone, tephra deposits were

sampled to guarantee rapid cooling of olivine phenocrysts, thus minimizing the diffusive

loss effects that occur during slow cooling in lava flows. Olivine crystals ranging in size

from 1 mm to 250 µm were hand-picked from washed and sieved tephra. Fifty olivine-

hosted melt inclusions were analyzed for H2O and δD by NanoSIMS at the Carnegie

Institution of Washington. Olivine was polished on a single side to expose the inclusions

(50-250 µm in diameter) and pressed into an indium mount along with basaltic glass

standard grains, and gold coated for analysis (40 nm thick coat). Each inclusion was

analyzed once with a 10 nA Cs+ primary beam (~2µm diameter) rastered over a 20x20

μm area. Ion images were used during pre-raster to verify homogeneity of the analyzed

area and avoid beam overlap with surrounding olivine and vapor bubbles (if present).

NanoSIMS analysis of D/H ratios in glass generally followed the methods and

standardization described in detail in Hauri et al. (2006) but with NanoSIMS instrument-

specific modifications described here. We used a modified entrance slit of 60 µm width

to maximize transmission of H and D ions, paired with a standard aperture slit (350 µm

width). This combination resulted in a mass resolving power of ~1000; previous

measurements44

have shown that with good vacuum conditions and when using a Cs+

Page 110: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

94

beam, H2/D ratios are typically <1.5x10-3

; this was verified at higher mass resolving

power prior to beginning analyses, with measured H2/D ≤1x10-4

. H and D were collected

on a Faraday Cup and electron multiplier, respectively; we also collected 18

O using a

Faraday cup. After presputtering and imaging to center the melt inclusion under the

beam without overlap onto olivine and/or vapor bubbles (where present), each isotope

was counted for one second per cycle, and we collected data over 300 total cycles for

each analysis. Faraday Cup backgrounds were measured off-peak 20 times every 80

ratios with the primary beam and electron beam on. Including presputtering, imaging and

measurement of signals and backgrounds, each analysis lasted approximately 15 minutes

and produced an average precision of ±3‰ (2M). Data were collected over 4 days of

analysis, and overall external reproducibility of the basalt glass standard GL07 D30-1

(1.5% H2O) was ±10‰ (2, n=32). D/H data were examined for matrix effects using a

two-collision cascade model described in detail in Hauri et al. (2006) but specific to the

NanoSIMS; the total range of matrix effects from hydrous rhyolite to low-H2O basalt on

the NanoSIMS are reduced (40‰) compared with the 6F ion microprobe (60‰), and

similarly it was found that for basaltic glass there are no resolvable matrix effects across

the range of H2O contents from 0.16 wt% to 2 wt% H2O (cf. Fig. 2.3 of Hauri et al.,

2006). Therefore we made no correction for matrix effects and simply corrected for the

instrumental mass fractionation (IMF) as determined by repeat analyses of GL07 D30-1

basaltic glass standard. The average precision of the IMF correction (±6‰ 2M of

standards run on each day, including uncertainty on known D/H ratio of GL07 D30-1)

was better than the overall external reproducibility; combining these uncertainties, we

estimate the typical uncertainty on a single SIMS measurement (precision + accuracy) is

Page 111: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

95

±12‰ (2). Simultaneous measurements of H2O content were determined from the

measured 1H/

18O ratio calibrated against the standard glasses described in Hauri et al.

(2006), with uncertainties of ±6% (2).

The same inclusions were subsequently analyzed for a suite of trace elements on

the Photon Machines Analyte G2 193 nm ArF ‘‘fast’’Excimer Laser system at Oregon

State University, using 50 µm spot size with a 5 Hz pulse rate. Measured trace element

concentrations were determined by reference to GSE-1G glass as a calibration standard

and using 43

Ca as an internal standard45

. BHVO-2G, BCR-2G, and GSD-1G glasses were

also analyzed to monitor accuracy and precision, and the analyzed values were within

10% of accepted values.

A.2. Supplementary Discussion

Composition of the hydrous subduction component

The curves in Figure 2.2a show the compositions of melts that can be formed by

partial melting of the mantle wedge after variable amounts of a hydrous subduction

component have been added. A hydrous subduction component (fluid/melt released from

the slab into the mantle wedge) was mixed with a 10% partial melt of a depleted MORB

mantle (DMM; Workman and Hart, 2005) and a more enriched mantle (E-M; 66% DMM

and 33% enriched mantle; Donnelly et al., 2004), which reflect the range in relative

mantle enrichment of sub-arc mantle seen in the Lassen region (Supplementary Fig. A.2).

For the subduction component, we have used an H2O/Ce ratio calculated for the central

Oregon Cascades (Ruscitto et al., 2010) using the method of Portnyagin et al. (2007), and

the value is consistent with the relatively high slab-surface temperature predicted from

Page 112: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

96

geodynamic models (Ruscitto et al., 2012). Because the concentrations of P2O5 in the

subduction component were not calculated by Ruscitto et al. (2010), we tried various

values, together with their maximum Sr concentration, until we achieved a good fit to the

data in Figure 2.2a. It is important to note that the absolute concentrations of elements in

subduction components calculated using the Portnyagin et al. (2007) method are poorly

constrained, as they result from summing the concentrations of all incompatible elements

(as oxides) plus H2O and Cl and then normalizing to 100%. The calculation does not

include SiO2 and many other major elements, so the concentrations of H2O and

incompatible trace elements would be quite different if the subduction component was a

hydrous silicate melt rather than an aqueous fluid. As a result, the mass fraction of the

subduction component added to the mantle along either of the curves in Figure 2.2a will

substantially vary depending on the fluid vs. melt-like character of the component.

However, the fraction of the total H2O in the final silicate melt generated in the wedge

that is contributed from the slab does not depend on this, and is a function only of the

initial mantle composition and the H2O/Ce content of the subduction component. For the

Cascades, the latter is well constrained from both geochemical data and geodynamic

models.

Calculation of slab surface temperatures

We use the H2O/Ce slab surface geothermometer (Cooper et al., 2012) for

comparison with temperatures calculated using a 2-D steady state thermal model (Wada

and Wang, 2009). The H2O/Ce ratio is a geochemical thermometer based on the

solubility of the REE-bearing phosphate minerals monazite and allanite (Cooper et al.,

Page 113: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

97

2012). These phases occur within the top-most sediment layers of the slab, making

H2O/Ce a potential thermometer of the slab surface. Basalt melting experiments at sub-

arc slab pressures and temperatures suggest that allanite is also stable in basaltic oceanic

crust (Kessel et al., 2005; Klimm et al., 2008; Plank et al., 2009), and that its presence in

experimental run products is not an artifact caused by REE doping of the experiments

(Klimm et al., 2008). Our model for wet melting of upper oceanic crust basalt provides a

mechanism for re-equilibration of the more deeply sourced, trace element-poor fluids

derived from serpentinized mantle in the downgoing plate, with an allanite-bearing

basaltic crust. Although there is uncertainty as to the melt fractions at which allanite is

exhausted from the residue during wet melting of basalt, the slab surface temperatures

calculated using the H2O/Ce ratio are in agreement with the thermal model temperatures.

If re-equilibration of the mantle-serpentinite-derived fluids with a REE-bearing phase did

not occur, then apparent H2O/Ce temperatures would be much lower because serpentinite

derived fluids are H2O rich and trace element poor. It should be noted, however, that our

model and interpretations are based on the slab surface temperatures derived from the 2-

D steady state thermal model and do not rely on the correctness of the H2O/Ce

temperatures.

In the H2O/Ce temperature calculation, we assume that fluid and/or melt

migration from the slab and into the mantle wedge are vertical. However, recent work has

shown that fluids may migrate updip at the slab-mantle interface (if fluids are mobile)

and within the slab (if a permeability barrier exists at the interface) and also down-dip

with the subducting slab or the flowing overlying mantle (if fluids are relatively

immobile). The details of fluid migration paths predicted by currently available fluid

Page 114: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

98

migration models (e.g., Wilson et al., 2014, Cagnioncle et al., 2007, Iwamori 1998, 2007)

vary widely, depending on the specific problems that are being investigated (e.g., some

combination of the effects of compaction, grain size, fluid availability, and melting) and

the assumptions and material properties used in the model. However, if lateral fluid

migration was taken into account, the effects would not change our general conclusions.

First, if up-dip flow was occurring (Wilson et al., 2014), then the fluids would be need to

be sourced from the slab down-dip of the arc. However, the thermal models predict that

no water remains in the slab past the arc. If metastability of hydrous minerals was

allowing water to be carried by the slab deeper than the arc, then the final stages of

chlorite breakdown would be occurring slightly down-dip of the arc, and the fluids rising

up-dip through the slab would have low D/H values. Second, in models that consider

lateral migration of fluids through the wedge once they are released from the slab

(Cagnioncle et al., 2007), the nature of the mantle wedge flow field controls the

migration. Initially, as fluids leave the slab-mantle wedge interface, they migrate in a

direction away from the arc trench because of down-dip flow of the wedge. But as the

fluids rise and trigger melting and the melts migrate through the upper part of the wedge,

the flow is directed back towards the trench because of the pattern of corner flow. Thus,

the two components of lateral migration partially offset each other. For most values of

wedge permeability, slab dip, and covergence velocity, there is a small net migration of

fluids and melts away from the trench (Cagnioncle et al., 2007), implying that arc

magmas will inherit a slab signature from a region of the slab that is slightly up-dip of the

region that lies directly beneath the arc. Overall then, the effects of some up-dip fluid

flow within the slab and lateral migration within the mantle wedge should partially offset

Page 115: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

99

each other. For these reasons, we have chosen the zeroth-order assumption that fluid flow

is simply vertical, and this is a good starting point from which further studies and

modeling efforts can depart.

For both models, we also choose to exclude the latent heat of fusion during slab

melting. Assuming a 10% partial melt of the slab, the latent heat of fusion may cool the

system by ~50C. However, the type of discrepancy between H2O/Ce and geodynamic

model temperatures that we find for the southern Cascades (Fig. 2.2) does not occur for

the central Oregon Cascades (Ruscitto et al., 2010, and our new unpublished modeling),

where a very similar thermal structure should also lead to slab melting. The small offset

between the model temperatures predicted for the southern Cascades and the lack of

offset for central Oregon is precisely what Cozzens et al. (2012) predicted should occur

based on fluid flow modeling and heat flow measurements. We thus think it is a better

explanation for the small discrepancy between H2O/Ce and geodynamic model

temperatures in the southern Cascades. Given that the effect of including the latent heat

of fusion is relatively small and that other secondary factors that could cause similar

small increases or decreases in model temperatures have yet to be included in

geodynamic models, we have chosen not to try to incorporate this into our modeling.

Hydrogen diffusive loss corrections

Supplementary Figure A.3 shows the H diffusion loss corrections made for each

sample using the diffusive re-equilibration model from Bucholz et al., (2013). Melt

inclusion data for each cinder cone are shown in a panel with a histogram and a bivariate

plot. Each histogram shows the H2O concentrations (wt%) measured by FTIR for a larger

Page 116: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

100

set of melt inclusions (Walowski and Wallace, unpublished data) than were analyzed by

NanoSIMS. Each panel also shows δD versus H2O (wt%) in melt inclusions measured by

NanoSIMS (Supplementary Fig. A.3). The curves show diffusion model results. For each

cone, the curves start at the maximum H2O concentration (defined as the average of the

highest 3-5 measured inclusions from the larger FTIR dataset), because these values are

assumed to be most representative of the starting, undegassed magma and are therefore

used as the starting value for the diffusive re-equilibration model. Model envelopes

represent variability in melt inclusion and host olivine size, both of which affect the

efficiency of diffusive loss. Starting δD for each model curve was varied in increments of

10‰ (colors shown in panel 1) to find model envelopes that best reproduced the

measured data points. The initial δD values shown in Figure 2.3a are the average of the

starting δD curves that fit the data, and uncertainties are reported as the upper and lower

limits of the starting δD values. Model corrections produce more negative values than are

measured at each cone. However, it is important to note that if the diffusion correction

was not performed and the lowest measured Lassen values were used to represent initial

δD, the values would still be distinctly more negative (isotopically lighter) than those

measured in the Marianas.

D/H Fractionation Model Calculations

The D/H fractionation model calculates the δD of the fluids released from the

downgoing slab using an equilibrium fractionation equation from White (2005),

11

(1 f ) ( f )

*1000

(Equation 1),

Page 117: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

101

where ∆ is the difference in the δD of the residual material (or solid) from the initial δD

of the starting material, f is the fraction of H2O remaining in the solid after the reaction,

and α is the fractionation factor. Once calculated, ∆ is used to then calculate the δD of the

residual slab (solid) as,

δDsolid = δDinitial +∆ (Equation 2).

Because the system is assumed to be in equilibrium, we can then use the definition of

alpha,

α = (δDsolid +1000) / (δDfluid +1000) (Equation 3),

to calculate the δD of the fluid as,

δDfluid = ((δDsolid +1000) / α) – 1000 (Equation 4).

The fraction of H2O remaining in the slab, f, was calculated by adopting the

approach developed by Wada et al., (2012). We use a generic lithologic model and initial

bulk H2O concentrations as described in Hacker (2008) and assume localized hydration in

the incoming oceanic plate (as opposed to uniform hydration; Wada et al., 2012). For the

Lassen Region, we use an incoming plate age of 8 Ma, a maximum depth of decoupling

of 75 km, and 2 km thick hydrated upper mantle, while in the Marianas case we assume a

4 km thick hydrated upper mantle (for other thermal model parameters, see Wada and

Wang, 2009, for the Cascades and Marianas). The modeling results for .the distribution

of H2O remaining in the slab are shown for the Lassen Region (Fig. 2.3b) and the

Marianas (Supplementary Fig. A.4). These results are consistent with calculations from

Van Keken et al., (2008). The starting δD values, δDinitial, were based on measured values

of natural samples for MORB and gabbro (-50‰; Agrinier et al., 1995) and serpentine in

Page 118: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

102

mantle peridotite (-60‰ to -40‰; Barnes et al., 2009; Alt and Shanks, 2006). The

fractionation factor, α, is calculated using a series of experimentally derived factors of

important hydrous phases found in the slab lithologies (Supplementary Fig. A.5).

Uncertainty in starting δD values (±10‰) and fractionation factors (Supplementary Fig.

A.5) result in an approximate error of ±15‰ (2σ) after propagation.

Previous work, dominantly in the Nicaraguan Arc, has provided evidence for the

hydration of the deep slab (oceanic peridotite up to ~20km depth; Ranero et al., 2005),

and the role of these fluids during the production of arc magmas (Barnes et al., 2009).

Furthermore, our model yields predictions that are very similar to values calculated for

fluids in equilibrium with exhumed chlorite-bearing harzburgites from Cerro del Almirez,

Spain, that have been interpreted to represent subducted oceanic peridotite. Alt et al.

(2012) calculated that fluids in equilibrium with these chlorite-bearing harzburgites at

700C had δD values that ranged from -65 to -27‰. This calculation employed the same

fractionation factors used in our study (Graham et al., 1984b, Graham at el. 1987), and

our methodologies only differ in that we model progressive dehydration. In our model for

the Cascades, 700C is reached in the hydrated upper mantle portion of the slab when the

slab surface is at ~60-65 km depth, at which point we predict δD of fluids released from

chlorite-bearing harzburgite to be -55 to -30‰, overlapping values from Alt et al. (2012).

However, because fractionation becomes more extreme as H2O concentrations decrease

in the peridotite at higher temperatures, large variations in δD in the released fluids occur

over a relatively small range in temperature. Our temperatures in the subducted mantle

beneath the arc are ~770C, producing fluids with δD values that range from

approximately -100 to -80‰, which overlap with Cascade melt inclusion values.

Page 119: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

103

Large-scale hydrogen isotope fractionation during fluid and/or melt transport

Hydrogen isotope fractionation could potentially occur by diffusion as fluids

and/or melts migrate through the slab and wedge. Indeed, this is the large-scale

equivalent of the process we observe and correct for in the melt inclusions (post-

entrapment H loss). But the significance of this process on the large scale is entirely

dependent on whether the H2O flux is dominantly diffusive or dominantly advective. We

postulate that veining within exhumed terrains that represent subducted oceanic

lithosphere (John et al., 2008; 2012) is evidence for dominantly advective fluid flow

within the slab and superjacent mantle wedge, and so, in our model we ignore diffusion

during large-scale dehydration of the slab. Fluid and melt transport within the mantle

wedge are also likely to be dominated by advective flow (Wilson et al., 2014, Cagnioncle

et al., 2007, Wada et al., 2012).

Sr isotope evidence for slab melting

In many arcs, basaltic magmas have variably elevated 87

Sr /86

Sr ratios that have

been interpreted to result from subduction recycling of seawater-derived Sr (e.g. DePaolo

and Wasserburg, 1977). However, in the southern Cascades, we observe negative

correlations between 87

Sr /86

Sr and subduction component indicators, opposite of what is

commonly observed in arcs (Supplementary Fig. A.6). The trend for the southern

Cascades requires a subduction component with unradiogenic 87

Sr /86

Sr (Borg et al.,

1997), which we suggest may be derived from partial melting of the downgoing basaltic

crust. This interpretation relies on the observation that oceanic crust is heterogeneously

Page 120: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

104

altered during hydrothermal alteration (e.g., Bach et al., 2003). Hydrous phases that form

during this alteration should acquire elevated 87

Sr /86

Sr values, whereas unaltered portions

should retain a MORB-like signature. Because hydrous phases are exhausted at the

eclogite transition, before the plate reaches subarc depths beneath the Cascades, much of

the radiogenic Sr that was added to the oceanic crust by hydration will likely be lost to

the forearc mantle wedge. Fluids from serpentine and chlorite dehydration within the slab

interior should have a seawater Sr-isotope signature, but are solute poor, and as a result

have little leverage when mixed with more Sr-rich, unradiogenic, partial melts of

eclogites (Spandler et al., 2014; Klimm et al., 2008).

Page 121: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

105

Supplementary Table A.1. Bulk Tephra Analyses

Sample BBL-05MI BORG-1 BPB-1 BAS-44-02 BPPC-01 BRVB-01 Cinder ConeMI

Lat (N) 40°34'33.72" 40°39'15.63" 40°40'40.80" 40°37'50.64" 40°19'34.22" 40°31'48.78" 40°32'24.50"

Long (W) 121°37'1.32" 121°13'29.07" 121°12'51.00" 121°20'40.14" 121°54'38.64" 121° 4'34.32" 121°18'37.00"

wt%

SiO2 49.76 49.29 52.44 50.04 51.83 50.38 51.85

TiO2 1.04 0.64 0.92 1.00 0.82 1.32 0.83

Al2O3 17.27 15.48 17.16 17.80 16.78 16.86 16.61

FeOt 9.38 8.54 7.65 8.88 7.68 9.39 6.88

MnO 0.14 0.15 0.14 0.16 0.15 0.16 0.09

MgO 8.38 8.05 8.76 9.01 9.14 8.65 9.68

CaO 10.34 8.50 9.35 9.92 10.60 9.24 10.02

Na2O 3.14 2.07 2.69 2.67 2.56 2.90 3.05

K2O 0.37 0.40 0.71 0.36 0.34 0.84 0.84

P2O5 0.18 0.10 0.19 0.16 0.10 0.27 0.15

CO2max

(ppm) 856 1199 614 721 1010 570 1138

H2Omax 1.33 3.30 3.23 1.17 2.44 2.67 3.46

Olivine Fo

content* 85.13 88.20 86.48 87.71 87.32 84.45 89.84

ppm

La 7.62 5.54 11.70 8.75 6.30 14.06 7.46

Ce 18.6 13.2 25.7 21.2 15.1 32.0 16.9

Pr 2.60 1.82 3.49 3.03 2.18 4.44 2.31

Nd 11.9 8.16 14.8 13.3 9.77 19.2 10.5

Sm 3.35 2.20 3.51 3.40 2.67 4.58 2.41

Eu 1.18 0.76 1.14 1.17 0.95 1.48 0.80

Gd 3.93 2.29 3.48 3.63 3.04 4.58 2.47

Dy 4.60 2.48 3.41 4.12 3.45 4.46 2.55

Er 2.81 1.39 1.88 2.46 2.12 2.42 1.39

Yb 2.66 1.31 1.68 2.27 1.95 2.12 1.39

Ba 146 154 309 151 125 337 202

Th 1.47 1.07 1.90 1.08 0.69 1.83 1.63

Nb 3.87 2.48 5.61 3.11 1.81 7.06 4.03

Y 24.6 12.9 17.5 22.0 18.9 22.9 13.1

Hf 2.63 1.61 2.53 2.33 1.83 2.91 1.73

Ta 0.25 0.17 0.37 0.20 0.12 0.45 0.26

U 0.44 0.33 0.57 0.34 0.23 0.55 0.52

Pb 3.32 2.48 3.84 2.66 1.89 3.44 3.22

Rb 8.68 7.40 14.60 7.24 6.25 16.03 15.39

Sr 186 362 464 411 461 459 342

Sc 41.4 28.8 30.2 36.7 36.0 30.1 27.1

Zr 104 60.1 98.5 89.3 67.2 114 70.9

MI – Values represent average of the most primitive measured melt inclusion compositions because bulk tephra

analyses were not available

*Average Fo content of all host-olivine of analyzed melt inclusions

Sample names are abbreviations based on Clynne and Muffler (2010): BBL = Basalt of Big Lake; BORG = Basalt of

Old Railroad Grade 3; BPB = Basalt of Poison Butte 3; BAS-44 = Basalt of Highway 44; and unpublished locations:

BPPC = Basalt of Paynes Creek Parasitic Cone; BRVB = Basalt of Round Valley Butte.

Locations are based on NAD27 datum used in Clynne and Muffler (2010). Additional geochemical data for most

samples can be found in Clynne et al. (2008); Borg et al. (1997, 2002, 2000); Clynne (1993).

Page 122: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

106

Supplementary Figure A.1. Bulk tephra trace element compositions. Trace element

compositions from bulk tephra of analyzed cones in the Lassen Region compared to

previously published endmembers. Low and high Sr/P represent endmember CABs with

variable amounts of a subduction component (Borg et al., 1997). The high-alumina

olivine tholeiite (HAOT, sometimes called low-K tholeiites; Bacon et al., 1997) is an

endmember composition of tholeiitic magmas from the Lassen region, associated with

normal faulting in the back-arc, which are not influenced by subduction enrichment.

Although some magmas used in this study have low K2O, they are more enriched with

respect to incompatible trace elements than the HAOT magmas, indicating that they are

enriched by a subduction component.

Page 123: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

107

Supplementary Figure A.2. Relative source depletion and subduction enrichment

Nb/Zr versus H2Omax/Ce showing the relative depletion of the source mantle (N-MORB

and E-MORB; Sun and McDonough, 1989; Michael, 1995) of the samples and effects of

subduction enrichment. The relative depletion of the mantle source seen in this figure is

used as the basis for calculating a range in starting mantle components used in Fig. 2.2a

and discussed in the Supplementary Materials. This figure also highlights that increased

addition of a slab component results in higher H2O/Ce (and therefore (Sr/P)N) and in

conjunction, creates source variability and across-arc trends seen in the Lassen region

primitive basalts.

Page 124: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

108

Supplementary Figure A.3. Corrections for effects of diffusive loss from melt

inclusions for each cone using the model of Bucholz et al., (2013). A panel for each

cinder cone contains a histogram (H2O concentrations measured by FTIR from a larger

set of inclusions) and a bivariate plot of δD versus H2O. Open circles are individual melt

inclusions measured by NanoSIMS, and colored curves are diffusion model results (color

based on starting δD shown in panel 1). The starting point for each model, H2Omax,

represents the average of 3-5 melt inclusions with the highest H2O measured by FTIR.

Page 125: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

109

Supplementary Figure A.4. a) Model results showing the distribution of H2O remaining

in the downgoing slab for the Marianas Arc and b) the cumulative H2O released from the

slab for both the Marianas and the Cascade Arc. Hydration of incoming plate is based on

those previously used in Wada et al., (2012).

Page 126: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

110

Supplementary Figure A.5. Experimentally derived fractionation factors (1000ln)

versus temperature for minerals found in a) MORB and gabbro and b) serpentinized

mantle. Fractionation factors used are for hornblende (Suzuoki and Epstein, 1976;

Graham et al., 1984a), epidote (Graham et al., 1980; Chacko et al., 1999), chlorite

(Graham et. al., 1984b), and serpentinite (Sakai and Tatsumi, 1978). In Fig. 5a, large

overlaps are seen between experimental results of similar phases, and thus individual

experimental data points are plotted along with best fit curves for each set of

experiments. The dashed line represents a best fit polynomial regression for all of the

data. Because of the range in fractionation factors for similar phases, we use the

regression that represents an average rather than calculating modal abundances of

minerals present in each lithology. The 2σ curves were used to calculate the total

uncertainty in the model contributed by the fractionation factors. For Fig. 5b, serpentinite

and chlorite are the only two phases for which fractionation factors exists, and the

temperature range over which experiments were performed do not overlap. Therefore, the

model uses serpentinite for T<500C and chlorite for T>500C.

Page 127: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

111

Supplementary Figure A.6. Sr isotopes versus mantle normalized Sr/P for basalts in

different Cascade Regions. Variation in 87

Sr/86

Sr in the Mt. Adams region (Jicha et al.,

2009) does not show negative correlations seen in the southern Cascades, likely due to a

distinct difference in mantle source composition, evidenced by abundant OIB-type

magmas in the southern Washington Cascades.

Page 128: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

112

Supplementary Figure A.7. Garnet signature in hot-slab subduction zones H2O/Ce

versus La/Yb from the global database of Ruscitto et al., (2012). Lassen Region data

points are averages of each cinder cone analyzed in this study.

Page 129: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

113

A.3. Hydrogen Isotope and Trace Element Compositions

The following table contains analyzed values for individual melt inclusions. H2O

wt% and δD were measured by NanoSIMS, and trace elements weremeasured by LA-

ICP-MS, as described in section A.1.

The range in PEC* listed in the table gives the calculated range in post-

entrapment crystallization (PEC) from a larger dataset of melt inclusions from the same

cinder cones (Chapter III). Major element compositions for the melt inclusions and host

olivine were not determined out for these inclusions, and we are therefore unable to

correct for post-entrapment modification. However, in the manuscript, we focus on

volatile and trace element ratios which should not be affected by PEC.

Page 130: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

114

BORG S-17 S-13 S-12 S-15 S-09 S-10 S-05 S-06 S-07 S-08 S-01 S-04 S-03

δD (‰) -11.0 -31.4 -47.9 -29.1 -58.8 -38.7 -42.0 -53.5 -25.1 -26.4 -63.4 -66.8 -48.8

H2O

(wt%) 2.37 2.93 2.51 2.61 2.71 2.54 2.62 2.85 2.44 2.67 3.01 2.92 2.94

Range in

% PEC*

0.20 - 7.5

(ppm)

Li 9.49 6.11 5.19 5.78 4.77 4.61 10.91 10.03 9.65 7.43 5.91 5.64 8.02

B 4.52 3.52 2.95 3.12 4.41 3.13 5.21 5.54 7.28 2.85 1.70 4.78 3.64

P 819 660 710 739 584 605 1291 1096 1009 726 320 1257 685

Ti 5633 4679 4695 4463 4016 4176 4081 6019 6334 4771 1550 7092 4562

V 236 236 318 237 242 262 348 289 288 249 164 307 257

Cu 305 78 110 174 104 103 152 129 369 110 96 138 112

Rb 8.20 5.36 6.03 5.10 4.72 4.95 7.95 7.84 9.78 5.88 1.53 10.25 5.10

Sr 660 533 487 353 410 411 368 602 707 523 541 635 422

Y 15.08 13.03 12.95 13.13 11.68 11.72 12.93 16.00 14.95 13.66 4.60 16.79 12.14

Zr 59.68 53.91 49.88 59.87 46.65 44.87 57.34 65.80 67.88 53.66 14.94 81.60 50.48

Nb 1.78 2.20 2.24 2.33 1.85 1.82 1.59 3.40 2.07 2.15 0.58 3.74 1.93

Ba 230 179 178 125 130 125 138 228 285 177 54 262 135

La 6.65 5.65 4.64 4.46 4.25 4.51 5.09 6.69 7.88 5.39 1.70 8.20 5.07

Ce 15.48 13.05 13.14 11.79 10.73 11.01 13.93 17.01 18.02 13.73 3.94 20.16 13.35

Pr 1.99 1.87 1.77 1.50 1.57 1.56 1.64 2.54 2.41 1.73 0.52 2.96 1.64

Nd 10.23 9.10 9.36 8.80 7.14 7.61 7.84 11.25 12.65 8.54 2.49 12.19 9.29

Sm 2.65 2.51 2.04 1.84 1.75 1.68 1.33 2.67 1.99 2.53 0.44 2.09 2.27

Eu 1.05 0.82 0.79 0.82 0.61 0.73 0.33 0.86 1.12 0.73 0.27 1.17 0.85

Gd 2.53 2.16 1.94 2.34 2.03 2.37 3.77 3.10 3.10 2.76 0.53 2.90 2.71

Dy 2.16 2.15 2.23 2.41 2.26 2.09 2.18 2.91 2.93 2.71 0.51 2.89 2.10

Er 1.43 1.33 1.47 1.38 1.35 1.41 N/A 1.65 1.69 1.73 0.50 1.91 1.32

Yb 1.75 1.63 1.65 1.36 1.22 1.24 N/A 1.67 1.42 1.03 0.47 1.60 1.45

Hf 1.95 1.28 1.35 1.26 1.23 1.09 1.10 1.21 1.78 1.31 0.32 1.58 1.50

Ta 0.07 0.12 0.12 0.11 0.05 0.09 N/A 0.13 N/A 0.05 N/A 0.14 0.07

Pb 4.28 2.90 3.42 1.71 2.38 2.50 2.27 3.46 5.17 2.86 0.82 5.09 3.23

Th 0.93 0.87 0.48 0.68 0.56 0.59 0.80 0.92 1.30 0.69 0.26 1.05 0.60

U 0.37 0.15 0.29 0.21 0.22 0.14 0.38 0.30 0.49 0.08 0.05 0.49 0.26

Page 131: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

115

BBL BBL-S-

06

BBL-S-

11

BBL-S-

08

BBL-S-

05

BBL-S-

04

BBL-S-

02

δD (‰) 5.17 -63.49 -59.52 -60.22 -79.41 -39.71

H2O

(wt%) 0.905 1.291 1.304 1.194 1.318 1.268

Range in

% PEC* 2.2-6.6

(ppm)

Li 6.59 5.75 5.04 6.05 10.25 12.15

B 1.38 0.63 1.38 6.23 2.64 N/A

P 874 942 973 3384 857 1237

Ti 5299 6266 6196 13031 5464 7100

V 206 202 219 310 217 259

Cu 151 58 83 75 71 114

Rb 3.67 4.85 5.23 16.51 3.54 7.20

Sr 226 262 256 387 251 266

Y 19.22 21.84 20.74 32.97 19.29 25.20

Zr 60.52 79.48 74.09 194.35 71.64 97.29

Nb 1.77 2.60 2.48 9.24 1.93 3.35

Ba 105 146 146 428 134 189

La 4.05 6.04 5.81 17.44 5.63 7.35

Ce 11.16 14.42 16.08 39.22 15.18 20.04

Pr 1.60 2.05 2.14 5.25 1.94 2.71

Nd 7.90 10.46 8.66 24.34 9.68 11.94

Sm 2.23 3.39 2.89 5.39 2.62 3.25

Eu 0.82 1.04 1.00 1.63 1.02 1.09

Gd 2.13 3.54 2.97 5.72 3.15 3.79

Dy 3.44 3.53 3.33 6.09 3.21 4.08

Er 2.00 2.42 3.00 3.80 2.37 2.41

Yb 2.14 2.50 3.27 3.50 1.77 3.26

Hf 1.51 1.68 1.97 4.90 1.45 2.30

Ta 0.10 0.16 0.09 0.61 0.13 0.20

Pb 1.85 1.85 1.92 6.02 1.86 3.58

Th 0.33 0.44 0.46 1.72 0.20 0.76

U 0.07 0.14 0.16 0.64 0.21 0.21

Page 132: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

116

CC S-09-

1

S-09-

5

S-09-

6

S-09-

7

S-09-

9

S-02-

6

S-02-

10

S-02-

3

S-02-

1

S-02-

8

δD (‰)

-

57.67

-

37.78

-

46.69

-

61.28

-

22.13

-

82.82 -50.95

-

71.94

-

60.02

-

55.45

H2O

(wt%) 2.554 2.680 2.734 2.556 2.654 2.760 2.559 2.627 2.421 2.524

Range in

% PEC* 3.1- 12.3

(ppm)

Li 8.85 11.27 11.34 11.48 9.33 15.27 12.72 10.35 10.66 11.02

B 9.23 7.76 6.41 4.95 5.61 6.41 6.53 5.68 6.17 7.20

P 1050 1123 1064 1068 1043 1732 1622 1160 1371 1856

Ti 4998 4777 4833 4701 4903 5972 6069 5802 5745 6153

V 214 220 204 239 218 236 224 191 208 260

Cu 75 91 84 82 77 86 97 69 72 95

Rb 18.16 17.35 18.56 16.91 16.23 23.04 19.94 16.86 16.87 19.40

Sr 376 344 383 355 357 419 431 424 412 405

Y 11.92 15.21 13.50 13.36 13.60 16.51 15.35 14.43 16.16 15.86

Zr 78.80 76.91 75.64 69.07 74.46 96.86 94.69 95.01 98.97 91.08

Nb 4.24 4.01 4.43 4.01 4.12 6.46 6.99 6.05 6.44 6.40

Ba 234 204 242 203 219 271 265 268 269 248

La 7.59 7.59 7.67 7.23 7.57 10.30 10.07 10.47 10.42 9.94

Ce 18.26 18.04 18.74 18.43 17.72 26.03 22.62 22.94 23.80 27.27

Pr 2.49 2.35 2.30 2.20 2.27 2.88 3.05 2.85 3.14 3.09

Nd 11.01 10.44 10.46 9.94 9.60 13.45 12.91 13.46 13.84 11.42

Sm 2.48 2.84 2.17 1.99 2.46 2.55 2.99 3.06 2.86 2.18

Eu 0.84 0.66 0.77 0.72 1.02 0.72 0.95 1.05 1.01 0.98

Gd 3.00 2.62 2.42 2.49 1.76 2.79 2.73 3.56 1.93 2.64

Dy 2.60 2.44 2.62 2.67 2.44 2.47 2.73 3.18 2.90 2.75

Er 1.56 1.78 1.26 1.95 1.40 2.12 1.83 2.09 1.82 1.89

Yb 1.40 1.34 1.58 1.55 1.68 2.11 1.40 2.57 1.50 1.72

Hf 1.91 1.86 2.09 1.61 1.93 2.24 1.99 2.12 2.41 2.05

Ta 0.12 0.25 0.17 0.19 0.22 0.28 0.41 0.39 0.46 0.30

Pb 3.65 3.54 3.72 2.87 3.15 3.80 3.92 3.36 3.89 6.31

Th 1.86 1.56 1.90 1.36 1.29 1.76 1.89 1.80 2.06 1.51

U 0.49 0.47 0.59 0.37 0.47 0.56 0.57 0.46 0.61 0.83

Page 133: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

117

BRVB BRVB-

S-14

BRVB-

S-11

BRVB-

S-9

BRVB-

S-8

BRVB-

S-13

BRVB-

S-6

BRVB-

S-1

δD (‰) -42.09 46.15 8.06 -23.27 4.63 162.27 82.13

H2O

(wt%) 2.187 1.632 1.922 2.236 1.918 0.542 1.372

Range in

% PEC* 3.1-10.2

(ppm)

Li 6.90 10.48 13.58 9.72 7.14 7.00 8.41

B 4.49 5.46 7.06 4.02 3.61 4.47 4.97

P 1891 1946 2269 1436 1566 1541 1623

Ti 7375 7300 7474 7241 7592 7072 7378

V 216 251 246 208 209 210 203

Cu 65 78 75 60 66 42 61

Rb 18.05 19.01 21.63 14.82 16.26 15.24 16.75

Sr 427 417 391 432 427 398 420

Y 17.15 18.04 19.67 18.14 17.18 19.37 18.24

Zr 96.59 96.99 109.14 110.82 105.87 94.69 104.66

Nb 6.61 6.78 7.04 5.98 6.45 6.08 6.30

Ba 337 335 327 361 350 311 333

La 12.91 12.22 13.79 13.46 13.79 12.79 13.08

Ce 29.56 29.42 32.24 32.64 32.20 27.83 29.09

Pr 3.51 3.71 3.77 3.96 3.91 3.88 4.08

Nd 17.23 16.13 18.26 20.70 17.47 16.12 16.80

Sm 4.05 4.04 3.55 4.47 3.63 4.08 3.80

Eu 1.26 1.19 1.14 1.43 0.92 1.19 1.16

Gd 3.66 3.50 4.30 3.68 4.05 3.43 3.56

Dy 3.34 3.24 3.68 4.20 3.61 3.76 3.52

Er 2.21 1.72 2.05 2.27 1.78 2.13 1.90

Yb 1.60 1.86 1.92 2.20 1.54 2.03 1.56

Hf 2.40 2.22 2.26 3.04 2.18 2.30 2.19

Ta 0.40 0.33 0.27 0.32 0.33 0.31 0.33

Pb 3.89 3.76 4.54 4.06 4.14 3.05 3.24

Th 1.31 1.55 1.71 2.07 1.66 1.34 1.62

U 0.49 0.56 0.40 0.40 0.53 0.41 0.51

Page 134: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

118

BPPC BPPC-

S-01

BPPC-

S-07

BPPC-

S-11

BPPC-

S-05

BPPC-

S-04

δD (‰) -47.21 -35.26 -45.98 -56.15 -46.61

H2O

(wt%) 2.127 2.161 2.154 2.194 2.206

Range in

% PEC* 0.3-12

(ppm)

Li 6.28 7.97 9.33 5.27 7.81

B 1.79 2.80 4.77 2.40 3.82

P 762 738 746 694 744

Ti 4602 4993 4929 4775 5100

V 246 279 231 232 249

Cu 296 14 28 38 32

Rb 4.49 7.14 6.00 5.89 6.94

Sr 451 470 469 448 455

Y 13.86 16.00 16.60 15.42 15.22

Zr 51.76 61.96 64.83 57.14 60.58

Nb 1.06 1.59 1.74 1.45 2.02

Ba 97 128 125 117 131

La 4.72 5.58 5.86 5.42 5.58

Ce 13.03 14.96 15.25 13.72 15.47

Pr 1.82 1.80 2.12 2.02 1.95

Nd 8.72 10.91 9.55 8.14 9.15

Sm 1.55 2.20 2.50 2.85 2.14

Eu 0.84 0.90 1.08 0.88 0.84

Gd 2.40 2.20 3.16 3.26 2.88

Dy 3.01 3.19 3.69 2.18 2.43

Er 1.74 2.30 2.37 1.95 1.44

Yb 1.40 1.32 1.72 1.66 1.69

Hf 1.50 1.10 1.42 1.56 1.35

Ta N/A N/A 0.07 0.10 0.14

Pb 3.22 3.40 2.04 2.07 2.65

Th 0.37 0.68 0.64 0.52 0.58

U 0.17 0.19 0.13 0.16 0.16

Page 135: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

119

BAS-44-

S-03

BAS-44-

S-01

BAS-44-

S-06

BPB-S-

02

BPB-S-

04

BPB-S-

07

BPB-S-

08

δD (‰) -52.65 -52.37 -29.73 -25.14 -29.12 -37.31 -34.92

H2O

(wt%) 1.135 1.169 0.863 2.299 2.485 2.576 2.506

Range in 1.8-4.3 (BAS-44)

0.1-12 (BPB)

% PEC*

(ppm)

Li 5.13 6.33 3.25 8.67 10.05 9.97 11.06

B 3.01 6.24 2.52 6.42 9.13 4.57 16.96

P 1127 949 1091 2017 1536 1356 1897

Ti 6353 6104 6185 6720 5218 5218 5328

V 229 223 232 255 230 212 256

Cu 50 127 133 97 84 77 100

Rb 5.03 3.76 6.20 16.46 16.38 12.96 25.72

Sr 440 389 416 420 450 450 465

Y 19.18 19.14 19.78 16.14 13.39 14.15 14.47

Zr 78.63 78.49 77.06 96.31 90.32 80.91 83.56

Nb 3.21 3.38 3.16 7.93 5.12 5.28 5.30

Ba 124 126 145 362 314 316 316

La 7.18 7.36 7.66 12.83 11.42 10.84 10.42

Ce 21.35 23.03 21.03 28.91 26.99 25.48 25.74

Pr 2.87 3.13 2.95 3.71 3.23 3.29 3.44

Nd 13.50 13.26 12.72 15.62 14.31 12.72 13.45

Sm 3.36 2.75 3.12 2.98 3.20 2.47 3.68

Eu 1.26 1.00 0.89 1.11 0.93 0.78 0.89

Gd 2.86 3.68 2.57 3.06 2.77 2.70 2.38

Dy 3.71 3.11 3.84 3.00 2.31 3.64 2.30

Er 2.59 2.47 2.42 1.44 1.49 1.86 1.49

Yb 1.64 1.09 2.09 0.84 1.76 1.41 1.37

Hf 1.57 1.78 1.69 2.95 2.30 2.29 1.91

Ta N/A N/A 0.25 0.25 0.33 0.26 0.30

Pb 1.92 1.59 2.93 4.31 5.25 3.70 6.15

Th 0.55 0.40 0.83 1.37 1.52 1.43 1.70

U 0.16 0.23 0.28 0.53 0.81 0.45 0.85

Page 136: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

120

Average Error

NanoSIMS

δD (‰) ±12‰ (2σ)

H2O (wt%) 0.12 wt% (2σ)

LA-ICP-MS

(ppm) 1 S.E.

Li 0.94

B 0.89

P 73

Ti 139

V 7.22

Cu 6.46

Rb 0.63

Sr 8.67

Y 0.53

Zr 2.41

Nb 0.26

Ba 4.9

La 0.35

Ce 0.63

Pr 0.15

Nd 0.64

Sm 0.32

Eu 0.11

Gd 0.38

Dy 0.28

Er 0.24

Yb 0.27

Hf 0.27

Ta 0.09

Pb 0.39

Th 0.15

U 0.16

Page 137: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

121

Supplementary References

Agrinier, P., Hékinian, R., Bideau, D., & Javoy, M. (1995). O and H stable isotope

compositions of oceanic crust and upper mantle rocks exposed in the Hess Deep

near the Galapagos Triple Junction. Earth and Planetary Science Letters, 136(3),

183-196.

Alt, J. C., & Shanks III, W. C. (2006). Stable isotope compositions of serpentinite

seamounts in the Mariana forearc: Serpentinization processes, fluid sources and

sulfur metasomatism. Earth and Planetary Science Letters, 242(3), 272-285.

Bach, W., Peucker‐Ehrenbrink, B., Hart, S. R., & Blusztajn, J. S. (2003). Geochemistry

of hydrothermally altered oceanic crust: DSDP/ODP Hole 504B–Implications for

seawater‐crust exchange budgets and Sr‐and Pb‐isotopic evolution of the

mantle. Geochemistry, Geophysics, Geosystems, 4(3).

Bacon C.R., Bruggman P.E., Christiansen, R.L, Clynne, M.A., Donnelly-Nolan, J.M.,

Hildreth, W. (1997). Primitive magmas at five Cascade volcanic fields: melts

from hot, heterogeneous sub-arc mantle. Canadian Mineralogist, 35. 397–42.

Barnes, J. D., Paulick, H., Sharp, Z. D., Bach, W., & Beaudoin, G. (2009). Stable isotope

(δ18

O, δD, δ37

Cl) evidence for multiple fluid histories in mid-Atlantic abyssal

peridotites (ODP Leg 209). Lithos, 110(1), 83-94.

Barnes, J. D., Sharp, Z. D., Fischer, T. P., Hilton, D. R., & Carr, M. J. (2009). Chlorine

isotope variations along the Central American volcanic front and back

arc. Geochemistry, Geophysics, Geosystems, 10(11).

Borg, L. E., Brandon, A. D., Clynne, M. A., & Walker, R. J. (2000). Re–Os isotopic

systematics of primitive lavas from the Lassen region of the Cascade arc,

California. Earth and Planetary Science Letters, 177(3), 301-317.

Cagnioncle, A. M., Parmentier, E. M., & Elkins‐Tanton, L. T. (2007). Effect of solid flow

above a subducting slab on water distribution and melting at convergent plate

boundaries. Journal of Geophysical Research: Solid Earth (1978–2012), 112(B9).

Chacko, T., Riciputi, R., Cole, R., & Horita, J. (1999). A new technique for determining

equilibrium hydrogen isotope fractionation factors using the ion microprobe:

Application to the epidote-water system. Geochimica et cosmochimica

acta, 63(1), 1-10.

Clynne, M. A. (1993). Geologic studies of the Lassen volcanic center, Cascade Range,

California. University of California, Santa Cruz.

Page 138: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

122

Clynne, M.A., Muffler, L. J. P., Siems, D. F., Taggart Jr, J. E., & Bruggman, P.

(2008). Major and EDXRF trace element chemical analyses of volcanic rocks

from Lassen Volcanic National Park and vicinity, California. US Geological

Survey.

Clynne, M. A., & Muffler, L. J. P. (2010). Geologic map of Lassen Volcanic National

Park and vicinity, California. US Department of the Interior, US Geological

Survey.

DePaolo, D. J., & Wasserburg, G. J. (1977). The sources of island arcs as indicated by Nd

and Sr isotopic studies. Geophysical Research Letters, 4(10), 465-468.

Donnelly, K. E., Goldstein, S. L., Langmuir, C. H., & Spiegelman, M. Origin of enriched

ocean ridge basalts and implications for mantle dynamics. Earth and Planetary

Science Letters, 226(3), 347-366 (2004).

Gaetani, G. A., O’Leary, J. A., Shimizu, N., Bucholz, C. E., & Newville, M. (2012).

Rapid reequilibration of H2O and oxygen fugacity in olivine-hosted melt

inclusions. Geology, 40(10), 915–918.

Graham, C. M., Sheppard, S. M., & Heaton, T. H. (1980). Experimental hydrogen isotope

studies—I. Systematics of hydrogen isotope fractionation in the systems epidote-

H2O, zoisite-H2O and AlO (OH)-H2O. Geochimica et Cosmochimica Acta, 44(2),

353-364.

Graham, C. M., Harmon, R. S., Sheppard, M. F. (1984). Experimental hydrogen isotope

studies: hydrogen isotope exchange between amphibole and water. American

Mineralogist, 69, 128-138.

Hacker, B. R. (2008). H2O subduction beyond arcs. Geochemistry Geophysics

Geosystems, 9(3).

Iwamori, H. (1998). Transportation of H2O and melting in subduction zones. Earth and

Planetary Science Letters, 160(1), 65-80.

Iwamori, H. (2007). Transportation of H2O beneath the Japan arcs and its implications for

global water circulation. Chemical Geology, 239(3), 182-198.

Jicha, B. R., Hart, G. L., Johnson, C. M., Hildreth, W., Beard, B. L., Shirey, S. B., &

Valley, J. W. (2008). Isotopic and trace element constraints on the petrogenesis of

lavas from the Mount Adams volcanic field, Washington. Contributions to

Mineralogy and Petrology, 157(2), 189–207.

Page 139: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

123

John, T., Klemd, R., Gao, J., & Garbe-Schönberg, C. D. (2008). Trace-element

mobilization in slabs due to non-steady-state fluid–rock interaction: constraints

from an eclogite-facies transport vein in blueschist (Tianshan, China). Lithos,

103(1), 1-24.

John, T., Gussone, N., Podladchikov, Y. Y., Bebout, G. E., Dohmen, R., Halama, R., and

Seitz, H. M. (2012). Volcanic arcs fed by rapid pulsed fluid flow through

subducting slabs. Nature Geoscience, 5(7), 489-492.

Kessel, R., Schmidt, M. W., Ulmer, P., & Pettke, T. (2005). Trace element signature of

subduction-zone fluids, melts and supercritical liquids at 120–180 km depth.

Nature, 437(7059), 724-727.

Klimm, K., Blundy, J. D., & Green, T. H. (2008). Trace element partitioning and

accessory phase saturation during H2O-saturated melting of basalt with

implications for subduction zone chemical fluxes. Journal of Petrology, 49(3),

523-553.

Michael, P. (1995). Regionally distinctive sources of depleted MORB: Evidence from

trace elements and H2O. Earth and Planetary Science Letters, 131(3), 301-320.

Newman, S., Epstein, S., & Stolper, E. (1988). Water, carbon dioxide, and hydrogen

isotopes in glasses from the ca. 1340 AD eruption of the Mono Craters,

California: constraints on degassing phenomena and initial volatile

content. Journal of Volcanology and Geothermal Research, 35(1), 75-96.

Plank, T., Cooper, L. B., & Manning, C. E. (2009). Emerging geothermometers for

estimating slab surface temperatures. Nature Geoscience, 2(9), 611-615.

Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., & Khubunaya, S. (2007).

Constraints on mantle melting and composition and nature of slab components in

volcanic arcs from volatiles (H2O, S, Cl, F) and trace elements in melt inclusions

from the Kamchatka Arc. Earth and Planetary Science Letters, 255(1-2), 53–69.

Ranero, C. R., Morgan, J. P., McIntosh, K., & Reichert, C. (2003). Bending-related

faulting and mantle serpentinization at the Middle America trench. Nature,

425(6956), 367-373.

Ruscitto, D. M., Wallace, P. J., Cooper, L. B., & Plank, T. (2012). Global variations in

H2O/Ce: 2. Relationships to arc magma geochemistry and volatile fluxes.

Geochemistry Geophysics Geosystems, 13(3).

Sakai, H. & Tsutsumi, M. (1978). D/H fractionation factors between serpentine and water

at 100 to 500 C and 2000 bar water pressure, and the D/H ratios of natural

serpentines. Earth Planetary Science Letters, 40: 231-242.

Page 140: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

124

Suzuoki, T., & Epstein, S. (1976). Hydrogen isotope fractionation between OH-bearing

minerals and water. Geochimica et Cosmochimica Acta, 40(10), 1229-1240.

White, W. M. (2005). Stable isotope geochemistry. Geochemistry: An Online Text

Book.

Wilson, C. R., Spiegelman, M., van Keken, P. E., & Hacker, B. R. (2014). Fluid flow in

subduction zones: The role of solid rheology and compaction pressure. Earth and

Planetary Science Letters, 401, 261-274.

Workman, R. K., & Hart, S. R. Major and trace element composition of the depleted

MORB mantle (DMM). Earth and Planetary Science Letters, 231(1), 53-72

(2005).

Page 141: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

125

APPENDIX B

CHAPTER III SUPPLEMENTARY MATERIALS

B.1. Restored Melt Inclusion Compositions

Data in this table represents individual melt inclusion compositions that have been

corrected for Fe loss and post-entrapment crystallization (PEC). Major elements were

restored using Petrolog 3.1.1 (Danyushevsky and Plechov, 2011). Volatile and trace

elements were restored by the same percent as K2O because all are incompatible in the

olivine host. Major element standard deviations represent an average of those calculated

for each melt inclusion from a given sample based on three EPMA point analyses. Trace

element standard deviations are also an average from individual inclusions from a given

sample based on the methods of Loewen and Kent (2012).

Page 142: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

126

SAMPLE BBL BBL BBL BBL BBL BBL BBL BBL BBL BBL BBL

avg

s.d.

INC # 1 3 4 6 8 9 13 15 18 20 21

wt%

SiO2 49.89 49.09 48.35 49.33 50.74 48.85 48.30 49.88 49.09 49.79 49.09 0.11

TiO2 1.22 0.96 0.82 1.09 1.27 0.99 0.88 1.26 1.05 1.07 1.01 0.03

Al2O3 16.85 17.29 16.64 16.90 15.77 17.53 17.42 16.66 17.22 16.43 16.97 0.15

FeOT 9.42 9.42 9.42 9.42 9.42 9.42 9.42 9.42 9.42 9.42 9.42 0.11

MnO 0.16 0.15 0.09 0.15 0.17 0.13 0.12 0.15 0.14 0.14 0.13 0.01

MgO 7.20 8.15 10.05 7.74 6.84 7.86 9.06 7.27 7.60 8.05 8.48 0.06

CaO 10.30 10.61 10.86 10.35 9.54 10.40 10.14 9.05 10.06 10.01 9.89 0.03

Na2O 3.17 2.96 2.54 3.11 3.51 3.18 2.90 3.50 3.20 3.19 3.09 0.02

K2O 0.46 0.30 0.15 0.39 0.71 0.37 0.28 0.72 0.41 0.39 0.37 0.01

P2O5 0.21 0.15 0.10 0.18 0.27 0.17 0.15 0.29 0.19 0.18 0.16 0.01

S 0.11 0.11 0.08 0.11 0.12 0.11 0.10 0.12 0.12 0.11 0.11 0.003

Cl 0.024 0.014 0.010 0.017 0.024 0.023 0.011 0.025 0.017 0.022 0.015 0.004

CO2

(ppm) 897 784 437 848 929 866 549 1034 852 712 856 69

H2O 0.94 0.77 0.80 1.17 1.47 0.94 1.18 1.57 1.47 1.15 1.23 0.06

% PEC 1.20 4.54 14.39 2.38 0.60 2.40 7.81 1.80 1.31 3.87 6.36

Fo 83.7 85.3 87.8 84.7 83.0 85.0 86.7 84.0 84.5 85.2 85.9

ppm

Li N/A N/A 8.92 7.78 9.18 7.41 7.78 9.38 5.96 8.05 6.07 1.50

B N/A N/A 3.27 1.87 5.16 0.47 2.63 3.59 1.82 2.42 1.38 1.13

Sc N/A N/A 45.14 40.10 40.70 36.03 33.19 42.02 41.90 38.61 37.65 3.33

V N/A N/A 248 236 266 221 209 219 226 234 226 13

Rb N/A N/A 2.06 5.67 9.74 5.21 4.06 8.16 6.11 6.52 5.40 0.31

Sr N/A N/A 193 275 308 270 249 307 285 269 275 5

Y N/A N/A 19.41 21.35 24.52 20.05 19.09 24.86 23.85 21.60 21.25 1.73

Zr N/A N/A 56.34 78.59 118.55 73.80 70.39 106.75 92.64 79.53 81.12 6.63

Nb N/A N/A 1.08 2.97 5.41 2.92 2.70 4.96 3.41 3.23 2.73 0.35

Ba N/A N/A 70.06 170.44 297.04 147.14 122.48 247.66 174.24 168.02 154.92 4.36

La N/A N/A 2.94 6.44 10.98 5.81 4.92 10.44 7.55 6.62 6.10 0.19

Ce N/A N/A 8.81 16.81 29.05 14.96 13.12 25.37 17.57 16.95 15.93 0.43

Pr N/A N/A 1.26 2.37 3.85 2.04 1.88 3.40 2.54 2.33 2.04 0.10

Nd N/A N/A 6.77 11.83 17.27 10.93 9.02 15.23 11.62 11.07 10.35 0.32

Sm N/A N/A 2.37 2.93 4.24 2.51 2.53 3.59 2.72 3.02 2.94 0.14

Eu N/A N/A 0.83 1.09 1.43 0.98 0.98 1.22 1.03 1.17 1.00 0.08

Gd N/A N/A 3.45 3.59 3.87 3.29 2.92 4.47 4.36 3.62 3.24 0.19

Dy N/A N/A 3.29 3.74 4.66 3.91 3.53 4.08 4.14 3.83 3.68 0.15

Er N/A N/A 2.41 2.47 2.73 2.38 2.17 3.18 2.85 2.74 2.47 0.14

Yb N/A N/A 2.45 2.44 2.84 2.24 2.42 3.07 2.47 2.60 2.67 0.15

Hf N/A N/A 1.36 1.95 2.58 1.55 1.50 2.57 1.96 2.09 1.61 0.11

Ta N/A N/A 0.06 0.17 0.36 0.13 0.15 0.29 0.19 0.18 0.19 0.03

Pb N/A N/A 0.89 2.15 3.65 1.70 1.46 3.13 2.08 2.01 1.70 0.11

Th N/A N/A 0.18 0.59 0.99 0.57 0.33 0.89 0.55 0.56 0.49 0.05

U N/A N/A 0.08 0.23 0.39 0.15 0.13 0.34 0.18 0.20 0.17 0.02

Page 143: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

127

BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB BPB

avg

s.d.

INC # 1 3a 3b 4 5 9 10a 10b 14a 14b 17 18 19 20 21 22

wt%

SiO2 50.77 51.21 51.27 50.14 50.44 51.45 50.98 50.39 51.42 52.17 51.42 50.83 49.88 51.26 51.07 51.16 0.24

TiO2 0.90 0.90 0.91 0.86 0.88 0.90 0.87 0.88 0.87 0.85 0.92 0.89 0.85 0.93 0.96 0.87 0.01

Al2O3 17.19 17.07 17.20 16.67 17.73 17.01 15.86 16.68 16.18 15.88 17.16 16.65 17.64 16.98 18.27 16.63 0.07

FeOT 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 7.65 0.07

MnO 0.12 0.12 0.12 0.11 0.11 0.12 0.09 0.11 0.10 0.13 0.15 0.12 0.11 0.16 0.11 0.09 0.01

MgO 7.07 6.85 6.84 7.88 7.46 6.84 7.91 7.90 7.77 7.73 7.73 7.83 7.69 7.38 6.14 7.29 0.18

CaO 10.37 10.05 10.13 9.67 9.89 10.21 10.02 10.38 9.96 9.77 10.40 10.00 10.13 10.25 10.08 10.12 0.10

Na2O 2.61 2.83 2.80 2.91 2.91 2.87 2.93 2.82 2.73 2.88 2.71 2.60 2.55 3.22 2.86 3.08 0.06

K2O 0.77 0.82 0.84 0.81 0.76 0.82 0.75 0.75 0.76 0.82 1.01 0.80 0.78 1.02 0.87 0.78 0.02

P2O 0.25 0.24 0.23 0.24 0.24 0.27 0.24 0.24 0.24 0.23 0.25 0.24 0.25 0.25 0.26 0.25 0.01

S 0.11 0.09 0.09 0.11 0.11 0.09 0.12 0.11 0.11 0.09 0.06 0.12 0.11 0.09 0.09 0.11 0.003

Cl 0.060 0.061 0.061 0.058 0.050 0.059 0.063 0.057 0.058 0.059 0.059 0.058 0.059 0.069 0.061 0.058 0.001

CO2

(ppm) 629 370 285 847 581 310 1563 1061 1123 475 135 412 1385 106 231 660 83

H2O 2.42 2.37 2.12 3.18 2.04 1.99 2.81 2.34 2.45 2.00 0.72 2.51 2.58 0.96 1.85 2.20 0.16

%

PEC 2.66 0.55 0.00 2.93 2.24 3.67 10.70 6.78 4.79 3.23 3.61 3.93 4.18 2.69 0.19 4.09

Fo 86.1 85.8 85.8 87.7 87.0 85.8 87.7 87.7 87.3 87.3 87.4 87.4 87.2 87.1 84.4 86.8

ppm

Li 8.63 8.49 N/A 7.21 4.54 8.17 15.49 14.52 8.12 10.36 10.73 N/A N/A 12.39 N/A 8.82 1.97

B 4.97 10.13 N/A 17.00 3.09 8.71 5.50 5.34 4.70 7.52 9.85 N/A N/A 6.63 N/A 11.88 1.68

Sc 34.01 35.74 N/A 35.39 34.32 36.58 53.90 41.41 38.30 28.17 38.40 N/A N/A 38.90 N/A 36.59 1.51

V 244 207 N/A 217 171 212 318 254 227 217 227 N/A N/A 263 N/A 226 11

Rb 15.22 16.97 N/A 17.14 11.37 14.87 18.36 14.47 14.90 15.57 18.53 N/A N/A 21.09 N/A 13.95 0.95

Sr 531 446 N/A 499 481 481 510 523 498 466 476 N/A N/A 452 N/A 551 7

Y 19.08 16.31 N/A 17.45 16.55 15.69 17.24 16.13 15.41 13.99 17.71 N/A N/A 18.47 N/A 15.60 0.76

Zr 108 90 N/A 102 96 92 107 96 97 101 106 N/A N/A 106 N/A 95 4

Nb 6.48 5.54 N/A 5.97 5.37 5.66 5.80 5.22 5.47 5.44 6.11 N/A N/A 6.88 N/A 6.28 0.36

Ba 328 304 N/A 336 258 308 359 338 304 311 374 N/A N/A 401 N/A 319 5

La 13.71 10.27 N/A 12.72 11.59 11.54 11.77 11.52 10.52 11.99 12.75 N/A N/A 13.38 N/A 12.23 0.21

Ce 28.33 22.96 N/A 23.35 20.34 25.62 35.87 29.11 25.95 25.69 28.74 N/A N/A 30.62 N/A 29.65 0.55

Pr 4.06 3.36 N/A 3.77 3.14 2.96 4.05 3.01 3.40 3.22 3.40 N/A N/A 4.02 N/A 3.54 0.15

Nd 14.55 15.10 N/A 15.95 11.24 13.80 15.26 18.17 13.45 13.35 17.43 N/A N/A 18.56 N/A 16.10 0.57

Sm 3.59 3.89 N/A 3.38 3.04 3.47 2.42 2.37 2.45 3.94 3.81 N/A N/A 3.60 N/A 3.07 0.33

Eu 1.20 1.03 N/A 1.15 1.14 1.02 1.09 1.02 1.01 0.95 1.10 N/A N/A 1.19 N/A 1.14 0.09

Gd 3.80 3.60 N/A 3.30 2.91 2.80 2.68 3.90 3.22 3.41 3.02 N/A N/A 4.82 N/A 3.29 0.18

Dy 3.18 3.20 N/A 3.09 2.30 2.92 3.38 3.30 3.12 2.52 2.85 N/A N/A 3.70 N/A 3.26 0.19

Er 1.69 1.61 N/A 1.73 1.91 2.43 1.68 1.46 1.84 3.26 1.91 N/A N/A 1.86 N/A 1.99 0.07

Yb 2.02 2.06 N/A 1.64 1.50 2.41 1.89 1.69 1.18 1.63 1.75 N/A N/A 1.77 N/A 1.85 0.14

Hf 2.69 2.45 N/A 2.24 3.38 2.64 2.36 2.36 1.16 2.41 3.09 N/A N/A 2.79 N/A 1.70 0.15

Ta 0.33 0.40 N/A 0.37 0.33 0.33 0.32 0.28 0.47 0.27 0.32 N/A N/A 0.39 N/A 0.24 0.03

Pb 4.29 4.73 N/A 4.46 2.92 4.65 4.76 4.43 4.51 3.33 5.30 N/A N/A 5.97 N/A 3.81 0.21

Th 2.78 1.93 N/A 1.57 1.53 2.08 1.59 1.60 1.34 1.91 1.99 N/A N/A 2.03 N/A 1.60 0.13

U 0.61 0.55 N/A 0.40 0.39 0.52 0.53 0.54 0.52 0.64 0.63 N/A N/A 0.70 N/A 0.64 0.06

Page 144: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

128

BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM BRM

avg

s.d.

INC # 1 2 3 5 6 7 8 9 10 11 14a 14b 15 16 18

wt%

SiO2 49.33 50.01 50.93 49.72 52.70 49.58 48.74 49.63 48.55 50.89 50.28 50.39 49.99 49.72 50.22 0.29

TiO2 0.85 0.88 0.80 0.86 0.89 0.87 0.93 0.80 0.82 0.82 0.91 0.91 0.92 0.94 0.87 0.01

Al2O3 18.74 20.14 18.15 17.32 15.05 17.86 17.64 17.58 17.87 18.93 19.82 19.87 18.19 17.75 19.11 0.22

FeOT 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 8.14 0.06

MnO 0.09 0.11 0.12 0.09 0.09 0.10 0.09 0.09 0.08 0.12 0.12 0.12 0.08 0.06 0.11 0.01

MgO 7.31 5.61 5.90 7.59 7.68 7.73 7.26 7.72 7.87 5.70 5.59 5.58 7.40 8.46 5.72 0.28

CaO 9.73 8.92 8.38 9.43 7.65 9.95 9.64 9.74 9.68 8.82 8.61 8.63 9.73 9.71 8.23 0.14

Na2O 4.12 4.26 3.89 3.55 3.45 3.46 3.87 3.85 3.75 3.89 4.06 4.07 3.86 3.43 3.90 0.15

K2O 0.70 0.83 0.79 0.75 0.91 0.72 0.78 0.64 0.71 0.80 0.75 0.75 0.73 0.86 0.75 0.02

P2O5 0.31 0.31 0.28 0.32 0.32 0.30 0.34 0.27 0.28 0.27 0.25 0.25 0.28 0.31 0.24 0.02

S 0.20 0.17 0.17 0.20 0.08 0.21 0.21 0.22 0.23 0.16 0.16 0.16 0.20 0.21 0.16 0.001

Cl 0.24 0.22 0.21 0.20 0.18 0.23 0.21 0.25 0.25 0.21 0.23 0.23 0.25 0.24 0.23 0.001

CO2

(ppm) 0 0 166 193 378 321 599 251 306 141 0 0 0 0 270 20

H2O 0.49 0.57 2.36 2.03 2.97 1.14 2.43 1.39 2.11 1.43 1.35 1.14 0.53 0.45 2.57 0.08

% PEC 8.13 2.19 1.20 10.26 8.42 7.48 7.21 8.10 8.15 2.11 3.79 3.72 9.26 12.75 1.23

Fo 86.5 82.8 83.0 86.5 86.4 86.8 86.4 87.0 87.3 82.6 82.6 82.6 86.5 87.9 82.8

ppm

Li 5.33 N/A 11.50 6.27 11.71 N/A N/A N/A N/A 4.51 N/A N/A 4.86 N/A 6.68 1.22

B 3.02 N/A 2.74 4.65 2.74 N/A N/A N/A N/A 2.18 N/A N/A 3.71 N/A 3.22 2.30

Sc 25.93 N/A 22.28 26.34 24.27 N/A N/A N/A N/A 20.18 N/A N/A 22.34 N/A 19.55 2.11

V 223 N/A 181 202 205 N/A N/A N/A N/A 168 N/A N/A 219 N/A 154 8

Rb 7.30 N/A 11.75 9.97 13.74 N/A N/A N/A N/A 10.24 N/A N/A 8.81 N/A 11.19 0.46

Sr 1378 N/A 1173 1257 969 N/A N/A N/A N/A 1039 N/A N/A 1395 N/A 1167 7

Y 11.08 N/A 11.44 11.39 12.27 N/A N/A N/A N/A 9.42 N/A N/A 11.72 N/A 12.25 0.83

Zr 85.70 N/A 92.69 92.57 98.34 N/A N/A N/A N/A 79.07 N/A N/A 95.32 N/A 84.94 5.05

Nb 3.96 N/A 4.58 5.14 6.73 N/A N/A N/A N/A 6.63 N/A N/A 5.26 N/A 4.25 0.49

Ba 236 N/A 283 285 332 N/A N/A N/A N/A 263 N/A N/A 284 N/A 257 8

La 14.62 N/A 14.98 15.78 16.43 N/A N/A N/A N/A 13.53 N/A N/A 16.02 N/A 13.95 0.25

Ce 35.41 N/A 36.69 37.51 37.32 N/A N/A N/A N/A 32.29 N/A N/A 37.74 N/A 31.52 0.51

Pr 4.20 N/A 4.27 4.41 4.71 N/A N/A N/A N/A 4.09 N/A N/A 4.65 N/A 4.19 0.08

Nd 17.50 N/A 19.52 19.33 20.35 N/A N/A N/A N/A 17.00 N/A N/A 19.73 N/A 17.92 0.30

Sm 3.61 N/A 2.82 3.54 2.98 N/A N/A N/A N/A 3.10 N/A N/A 3.64 N/A 3.42 0.26

Eu 1.11 N/A 1.12 1.14 1.15 N/A N/A N/A N/A 1.09 N/A N/A 1.30 N/A 0.95 0.06

Gd 2.97 N/A 2.88 2.75 2.49 N/A N/A N/A N/A 2.87 N/A N/A 3.06 N/A 2.97 0.07

Dy 2.36 N/A 1.96 2.45 2.07 N/A N/A N/A N/A 1.54 N/A N/A 2.40 N/A 2.00 0.13

Er 1.44 N/A 1.31 1.39 1.16 N/A N/A N/A N/A 0.64 N/A N/A 1.36 N/A 1.39 0.16

Yb 1.34 N/A 1.27 1.77 1.11 N/A N/A N/A N/A 0.71 N/A N/A 1.23 N/A 1.44 0.20

Hf 2.03 N/A 2.78 2.35 2.32 N/A N/A N/A N/A 1.69 N/A N/A 2.58 N/A 2.54 0.12

Ta 0.17 N/A 0.22 0.24 0.29 N/A N/A N/A N/A 0.27 N/A N/A 0.23 N/A 0.18 0.04

Pb 4.57 N/A 5.08 5.08 5.80 N/A N/A N/A N/A 4.15 N/A N/A 4.39 N/A 4.63 0.17

Th 1.30 N/A 1.94 1.75 1.82 N/A N/A N/A N/A 1.46 N/A N/A 1.68 N/A 1.82 0.11

U 0.46 N/A 0.47 0.47 0.60 N/A N/A N/A N/A 0.38 N/A N/A 0.51 N/A 0.51 0.04

Page 145: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

129

BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB BRVB s.d.

INC# 1 3 4 5 6 8 9 10 11 13 18 20 21 22

wt%

SiO2 48.76 49.64 49.12 49.26 49.05 49.23 53.11 49.47 48.35 49.67 49.08 50.96 48.27 48.49 0.17

TiO2 1.29 1.47 1.36 1.34 1.34 1.45 2.37 1.36 1.35 1.31 1.33 1.42 1.44 1.32 0.03

Al2O3 17.81 15.97 16.28 17.50 16.42 16.92 14.18 15.51 16.57 15.78 16.95 15.41 18.48 16.43 0.08

FeOT 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 9.39 0.06

MnO 0.12 0.13 0.13 0.14 0.14 0.14 0.13 0.14 0.13 0.15 0.14 0.12 0.14 0.15 0.01

MgO 7.44 7.35 7.52 7.28 7.45 6.95 6.99 7.15 7.19 8.09 7.21 7.26 6.67 7.28 0.03

CaO 8.67 9.25 9.59 9.46 9.80 9.85 7.99 9.64 9.24 9.59 9.46 9.37 9.91 9.35 0.03

Na2O 3.75 3.27 3.31 3.38 3.32 3.77 2.91 3.27 3.60 3.54 3.49 3.34 3.67 3.24 0.07

K2O 1.25 1.02 1.01 1.05 0.96 1.13 1.66 0.98 1.04 0.93 0.94 1.18 1.13 0.95 0.02

P2O5 0.30 0.32 0.29 0.32 0.29 0.35 0.51 0.30 0.28 0.26 0.29 0.27 0.33 0.28 0.01

S 0.14 0.19 0.16 0.19 0.20 0.17 0.01 0.20 0.19 0.17 0.17 0.04 0.16 0.20 0.002

Cl 0.038 0.048 0.049 0.049 0.049 0.049 0.056 0.050 0.049 0.039 0.039 0.038 0.049 0.04 0.002

CO2

(ppm) 657 748 595 644 0 287 403 410 619 419 551 455 123 560 34

H2O 1.00 2.01 1.82 0.71 1.68 0.63 0.57 2.61 2.70 1.12 1.57 1.07 0.38 2.89 0.10

% PEC 6.85 6.87 4.79 3.55 3.87 3.16 6.53 2.01 2.57 4.77 2.88 7.67 3.23 2.78

Fo 85.0 84.4 84.8 84.4 84.7 84.2 83.5 84.0 84.5 85.9 84.3 84.2 83.6 84.2

ppm

Li N/A 10.77 10.13 N/A 10.19 N/A N/A 12.11 9.86 N/A 16.23 N/A 0.26 10.88 1.70

B N/A 6.17 16.19 N/A 5.28 N/A N/A 114.73 4.86 N/A 11.14 N/A 10.81 4.25 2.80

Sc N/A 36.50 30.06 N/A 31.16 N/A N/A 33.44 29.58 N/A 37.74 N/A 48.40 27.67 1.43

V N/A 262 230 N/A 249 N/A N/A 235 239 N/A 310 N/A 408 240 13

Rb N/A 21.73 18.91 N/A 19.68 N/A N/A 26.62 18.56 N/A 25.51 N/A 32.25 19.88 0.53

Sr N/A 456 457 N/A 453 N/A N/A 454 437 N/A 469 N/A 326 455 7

Y N/A 18.88 18.69 N/A 18.69 N/A N/A 17.31 17.81 N/A 19.29 N/A 33.83 17.25 0.74

Zr N/A 110 107 N/A 103 N/A N/A 102 99 N/A 112 N/A 200 99 5

Nb N/A 6.77 6.77 N/A 6.66 N/A N/A 7.43 6.77 N/A 7.63 N/A 11.80 6.91 0.33

Ba N/A 412 359 N/A 361 N/A N/A 386 347 N/A 425 N/A 548 366 7

La N/A 14.20 13.56 N/A 14.08 N/A N/A 13.98 13.20 N/A 16.37 N/A 23.82 13.67 0.22

Ce N/A 35.99 32.43 N/A 32.93 N/A N/A 33.25 32.71 N/A 43.15 N/A 51.57 33.63 0.70

Pr N/A 4.45 4.23 N/A 4.16 N/A N/A 4.32 4.07 N/A 5.15 N/A 7.43 4.22 0.10

Nd N/A 20.63 18.70 N/A 18.95 N/A N/A 20.17 18.27 N/A 20.61 N/A 34.49 20.31 0.43

Sm N/A 4.23 4.35 N/A 4.00 N/A N/A 4.38 4.00 N/A 4.67 N/A 7.10 4.16 0.13

Eu N/A 1.40 1.31 N/A 1.36 N/A N/A 1.31 1.38 N/A 1.45 N/A 1.99 1.25 0.08

Gd N/A 3.97 3.92 N/A 3.91 N/A N/A 3.88 3.73 N/A 4.68 N/A 6.88 3.61 0.21

Dy N/A 3.53 3.60 N/A 3.46 N/A N/A 3.57 3.37 N/A 4.28 N/A 6.02 3.82 0.20

Er N/A 2.12 1.91 N/A 2.05 N/A N/A 1.85 1.94 N/A 2.13 N/A 3.55 1.79 0.08

Yb N/A 2.00 1.67 N/A 1.92 N/A N/A 1.65 1.73 N/A 2.78 N/A 3.21 1.70 0.05

Hf N/A 2.69 2.56 N/A 2.74 N/A N/A 2.38 2.31 N/A 2.59 N/A 4.36 2.51 0.13

Ta N/A 0.38 0.38 N/A 0.32 N/A N/A 0.34 0.29 N/A 0.42 N/A 0.67 0.34 0.06

Pb N/A 4.11 3.85 N/A 3.84 N/A N/A 3.92 3.79 N/A 5.32 N/A 6.75 4.15 0.18

Th N/A 1.77 1.61 N/A 1.76 N/A N/A 1.73 1.70 N/A 1.99 N/A 2.67 1.80 0.12

U N/A 0.61 0.54 N/A 0.63 N/A N/A 0.63 0.61 N/A 0.65 N/A 1.00 0.62 0.05

Page 146: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

130

BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC BPPC s.d.

INC# 1 2 3 5 7 10 11 12 13 15 17a 17b 18 19

wt%

SiO2 50.52 50.21 50.13 49.64 50.62 50.17 50.27 50.68 51.03 51.55 50.39 50.31 50.22 50.44 0.29

TiO2 0.78 0.78 0.71 0.76 0.80 0.79 0.77 0.81 0.76 1.01 0.80 0.80 0.82 0.77 0.02

Al2O3 16.42 16.52 16.65 16.64 16.47 16.48 16.56 16.47 15.38 15.59 16.24 16.76 16.72 16.72 0.13

FeOT 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 7.68 0.07

MnO 0.13 0.11 0.10 0.12 0.12 0.11 0.13 0.11 0.13 0.18 0.11 0.12 0.14 0.14 0.01

MgO 7.88 8.40 8.43 8.26 7.98 8.04 8.06 8.13 7.55 8.20 8.15 8.19 7.75 7.75 0.18

CaO 10.79 10.71 11.06 11.52 10.77 11.01 10.77 10.89 11.34 11.34 10.93 10.69 11.12 10.89 0.10

Na2O 2.71 2.68 2.58 2.38 2.73 2.69 2.67 2.82 2.63 3.04 2.77 2.71 2.85 2.80 0.11

K2O 0.31 0.30 0.24 0.19 0.33 0.31 0.35 0.32 0.30 0.45 0.31 0.30 0.36 0.31 0.02

P2O5 0.18 0.17 0.18 0.16 0.19 0.19 0.19 0.18 0.21 0.22 0.19 0.18 0.19 0.18 0.01

S 0.12 0.11 0.11 0.13 0.12 0.12 0.12 0.12 0.12 0.07 0.11 0.11 0.13 0.11 0.003

Cl 0.039 0.039 0.037 0.040 0.040 0.039 0.040 0.039 0.030 0.050 0.039 0.038 0.040 0.040 0.001

CO2

(ppm) 908 878 954 758 1300 1008 1012 672 892 85 876 1011 954 735 52

H2O 2.50 2.32 2.12 2.54 2.19 2.43 2.44 1.79 2.85 0.61 2.32 2.15 2.05 2.21 0.11

% PEC 1.48 3.37 9.49 1.47 0.82 2.59 1.13 5.00 1.79 1.26 3.17 4.40 0.82 0.87

Fo 87 88 88 88 87 87 87 88 86 88 88 88 87 87

ppm

Li 6.14 6.06 6.89 N/A N/A 5.80 7.25 N/A 4.00 6.99 7.35 N/A 7.04 N/A 1.36

B 1.74 2.69 1.31 N/A N/A 1.32 2.30 N/A 4.61 3.43 N/A N/A 4.34 N/A 1.77

Sc 35.87 34.04 31.48 N/A N/A 33.30 33.74 N/A 35.34 40.37 36.71 N/A 36.15 N/A 1.45

V 211 212 215 N/A N/A 211 208 N/A 209 238 210 N/A 210 N/A 13

Rb 5.09 4.58 2.72 N/A N/A 4.46 4.95 N/A 5.50 6.75 4.74 N/A 6.16 N/A 0.50

Sr 455 447 345 N/A N/A 456 432 N/A 458 378 445 N/A 455 N/A 6

Y 14.99 15.36 13.39 N/A N/A 16.18 15.15 N/A 15.88 18.39 16.19 N/A 16.12 N/A 0.78

Zr 58.30 60.60 41.62 N/A N/A 54.75 58.19 N/A 58.42 68.02 59.08 N/A 61.90 N/A 3.93

Nb 1.41 1.42 0.83 N/A N/A 1.39 1.37 N/A 1.30 1.71 1.40 N/A 1.65 N/A 0.34

Ba 111 102 61 N/A N/A 102 105 N/A 119 136 105 N/A 118 N/A 4

La 5.14 5.23 3.79 N/A N/A 5.33 5.60 N/A 5.74 6.29 4.99 N/A 5.61 N/A 0.20

Ce 13.10 12.54 9.59 N/A N/A 12.52 13.01 N/A 12.91 16.22 13.36 N/A 13.57 N/A 0.36

Pr 1.88 1.66 1.23 N/A N/A 1.66 1.70 N/A 1.85 2.00 1.95 N/A 1.92 N/A 0.13

Nd 9.04 8.29 6.11 N/A N/A 8.78 8.74 N/A 8.48 9.52 9.26 N/A 8.53 N/A 0.24

Sm 2.37 2.56 1.87 N/A N/A 2.16 1.96 N/A 2.42 3.00 1.32 N/A 2.85 N/A 0.12

Eu 0.99 0.78 0.65 N/A N/A 0.75 0.83 N/A 0.82 0.85 0.87 N/A 0.92 N/A 0.06

Gd 2.48 2.62 1.77 N/A N/A 2.77 2.64 N/A 3.73 2.42 2.48 N/A 2.68 N/A 0.12

Dy 2.78 2.38 2.27 N/A N/A 2.75 2.73 N/A 2.74 3.16 2.88 N/A 2.43 N/A 0.08

Er 1.57 1.63 1.16 N/A N/A 1.88 1.62 N/A 1.71 1.91 2.47 N/A 1.76 N/A 0.18

Yb 1.70 1.56 1.54 N/A N/A 1.75 1.86 N/A 2.02 1.75 1.62 N/A 1.62 N/A 0.10

Hf 1.66 1.52 1.16 N/A N/A 1.61 1.44 N/A 1.90 1.51 1.95 N/A 1.88 N/A 0.13

Ta 0.09 0.06 N/A N/A N/A 0.05 0.13 N/A 0.04 0.12 0.07 N/A 0.10 N/A 0.03

Pb 1.64 1.56 1.02 N/A N/A 1.50 1.51 N/A 1.96 2.33 1.39 N/A 1.76 N/A 0.18

Th 0.54 0.48 0.30 N/A N/A 0.45 0.60 N/A 0.46 0.69 0.62 N/A 0.60 N/A 0.08

U 0.18 0.17 0.01 N/A N/A 0.16 0.19 N/A 0.25 0.22 0.15 N/A 0.19 N/A 0.03

Page 147: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

131

CC CC CC CC CC CC CC CC CC CC CC CC CC

avg

s.d.

INC# 1 2 6 7 9a 9b 4a 4b 10 13 16a 16b 18

wt%

SiO2 51.11 50.20 49.91 50.71 51.11 50.76 50.63 50.88 50.75 52.64 49.96 49.87 50.88 0.34

TiO2 0.87 0.94 0.86 0.83 0.81 0.80 0.85 0.83 0.81 0.81 0.83 0.84 0.83 0.03

Al2O3 16.65 17.98 16.55 16.65 16.27 16.49 16.82 16.59 16.82 16.71 16.47 16.53 16.44 0.27

FeOT 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 6.88 0.13

MnO 0.09 0.09 0.08 0.09 0.09 0.08 0.08 0.09 0.08 0.11 0.10 0.09 0.09 0.01

MgO 7.93 7.47 8.44 8.37 8.38 8.47 7.88 7.95 8.28 7.67 8.25 8.27 8.18 0.75

CaO 10.10 9.42 10.67 10.13 10.15 9.73 10.21 9.90 10.12 8.43 10.68 10.73 10.18 0.22

Na2O 3.24 3.43 2.89 2.77 3.02 2.85 3.17 3.03 3.02 3.52 2.97 2.93 3.07 0.16

K2O 0.86 0.84 0.72 0.82 0.79 0.82 0.81 0.88 0.82 1.27 0.68 0.68 0.79 0.03

P2O5 0.16 0.18 0.16 0.16 0.15 0.15 0.16 0.15 0.16 0.15 0.16 0.15 0.15 0.01

S 0.09 0.09 0.09 0.09 0.09 0.05 0.09 0.09 0.09 0.08 0.09 0.09 0.09 0.003

Cl 0.036 0.038 0.037 0.034 0.040 0.038 0.038 0.035 0.038 0.036 0.037 0.037 0.036 0.003

CO2

(ppm) 922 940 1302 1248 0 789 531 640 1151 612 1431 1464 1266 85

H2O 2.58 3.05 3.33 3.10 2.84 3.46 3.00 3.30 2.74 2.29 3.49 3.52 2.99 0.23

% PEC 6.11 1.70 8.02 0.81 7.49 7.68 6.06 6.76 7.42 6.46 6.05 6.33 6.49

Fo 90 89 90 90 90 90 90 90 90 90 90 90 90

ppm

Li N/A N/A 5.85 9.67 15.71 N/A 7.69 7.20 N/A 8.60 24.63 4.83 N/A 1.23

B N/A N/A 7.20 5.42 6.52 N/A 5.10 2.29 N/A 5.67 9.41 3.83 N/A 2.93

Sc N/A N/A 29.83 30.52 31.25 N/A 26.88 26.93 N/A 25.40 23.61 29.22 N/A 1.36

V N/A N/A 196 206 177 N/A 193 174 N/A 189 162 192 N/A 11

Rb N/A N/A 10.20 17.00 13.96 N/A 13.48 13.65 N/A 12.56 26.23 9.41 N/A 0.71

Sr N/A N/A 347 370 367 N/A 343 346 N/A 325 349 320 N/A 8

Y N/A N/A 15.38 14.10 15.63 N/A 13.02 13.19 N/A 12.48 12.73 13.24 N/A 0.86

Zr N/A N/A 80.98 75.69 89.69 N/A 68.49 71.84 N/A 68.56 73.03 67.73 N/A 4.29

Nb N/A N/A 4.46 4.34 4.05 N/A 3.95 4.04 N/A 3.80 3.90 4.16 N/A 0.33

Ba N/A N/A 178 205 206 N/A 189 198 N/A 182 280 157 N/A 5

La N/A N/A 8.38 7.81 8.05 N/A 7.12 7.54 N/A 6.96 8.77 6.57 N/A 0.47

Ce N/A N/A 17.97 17.59 18.42 N/A 16.72 16.82 N/A 16.13 18.84 15.44 N/A 0.77

Pr N/A N/A 2.50 2.31 2.74 N/A 2.28 2.54 N/A 2.23 2.32 2.17 N/A 0.08

Nd N/A N/A 10.96 10.34 15.03 N/A 10.61 11.35 N/A 10.11 11.53 9.20 N/A 0.37

Sm N/A N/A 2.22 2.61 2.06 N/A 2.30 2.47 N/A 2.51 2.31 2.26 N/A 0.22

Eu N/A N/A 1.03 0.90 1.08 N/A 0.81 0.79 N/A 0.70 0.79 0.79 N/A 0.09

Gd N/A N/A 3.04 2.74 2.87 N/A 2.54 2.75 N/A 2.17 2.33 2.30 N/A 0.24

Dy N/A N/A 2.63 2.56 2.65 N/A 2.30 3.03 N/A 2.58 2.35 2.45 N/A 0.11

Er N/A N/A 1.66 1.38 2.46 N/A 1.40 1.21 N/A 1.29 1.42 1.61 N/A 0.10

Yb N/A N/A 1.43 1.59 1.74 N/A 1.28 1.55 N/A 1.33 1.31 1.30 N/A 0.14

Hf N/A N/A 1.77 1.92 3.45 N/A 1.58 1.91 N/A 1.45 1.82 1.70 N/A 0.15

Ta N/A N/A 0.28 0.25 0.43 N/A 0.24 0.18 N/A 0.38 0.24 0.27 N/A 0.03

Pb N/A N/A 2.40 2.91 2.86 N/A 2.84 2.88 N/A 2.46 6.59 1.66 N/A 0.13

Th N/A N/A 1.29 1.64 1.65 N/A 1.55 1.65 N/A 1.53 2.27 1.12 N/A 0.08

U N/A N/A 0.35 0.51 0.54 N/A 0.46 0.58 N/A 0.46 0.76 0.37 N/A 0.04

Page 148: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

132

BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG BORG s.d.

INC # A1 A3 A5 A6 c A9 A10 A18

A20-

2 A23 A24

A24-

2 A28

AA1

c AA4 AA5

wt%

SiO2 47.78 48.41 48.65 49.61 48.20 48.45 49.08 49.31 48.70 48.31 48.46 48.41 49.93 49.15 49.42 0.41

TiO2 0.66 0.67 0.71 0.82 0.64 0.66 0.72 0.80 0.72 0.70 0.68 0.69 0.83 0.85 0.74 0.02

Al2O3 16.95 16.75 16.23 19.21 16.35 17.47 15.72 17.87 16.30 15.89 16.17 17.03 16.88 17.90 15.88 0.28

FeOT 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 8.54 0.11

MnO 0.12 0.11 0.12 0.12 0.11 0.10 0.09 0.13 0.10 0.08 0.10 0.10 0.12 0.12 0.09 0.01

MgO 9.80 9.73 9.51 6.97 9.77 9.69 9.47 7.44 8.99 10.32 10.28 9.67 8.14 7.90 9.34 0.14

CaO 9.76 9.96 9.97 8.73 10.02 9.03 10.13 9.16 10.02 10.73 10.62 10.38 8.76 8.71 9.99 0.11

Na2O 2.63 2.48 2.48 2.46 2.41 2.73 2.41 3.27 2.56 2.36 2.49 2.46 2.74 3.22 2.47 0.14

K2O 0.50 0.47 0.51 0.55 0.49 0.50 0.48 0.52 0.50 0.38 0.40 0.46 0.57 0.54 0.48 0.02

P2O5 0.10 0.07 0.09 0.13 0.07 0.08 0.10 0.10 0.10 0.09 0.09 0.08 0.11 0.11 0.11 0.01

S 0.10 0.11 0.12 0.15 0.12 0.12 0.12 0.11 0.12 0.11 0.12 0.12 0.10 0.21 0.11 0.004

Cl 0.05 0.05 0.05 0.05 0.05 0.06 0.05 0.04 0.05 0.06 0.06 0.05 0.04 0.04 0.05 0.002

CO2

(ppm) 518 853 1114 724 945 887 1493 841 1104 1123 950 793 786 840 984 53

H2O 3.00 2.67 3.07 2.74 3.27 2.62 3.11 2.73 3.34 2.45 2.04 2.04 3.26 2.75 2.80 0.12

PEC 9.80 10.13 9.27 9.25 9.20 19.03 11.61 3.15 13.96 13.95 12.61 14.58 8.99 5.27 12.74

Fo 88.6 88.4 88.2 83.8 88.5 88.4 88.0 85.4 87.5 89.1 89.1 88.4 86.2 86.1 87.9

ppm

Li 7.60 N/A 5.76 N/A 5.43 N/A 10.48 6.51 5.01 6.75 N/A 7.74 6.01 N/A 7.66 1.18

B 2.87 N/A 3.36 N/A 4.41 N/A 3.82 6.91 2.49 4.09 N/A 4.99 3.72 N/A 4.98 1.77

Sc 38.35 N/A 35.93 N/A 37.47 N/A 41.11 29.58 35.29 41.35 N/A 39.69 30.05 N/A 39.37 1.60

V 272 N/A 245 N/A 250 N/A 339 227 236 288 N/A 305 225 N/A 280 16

Rb 5.20 N/A 4.25 N/A 5.95 N/A 6.91 8.83 4.37 5.63 N/A 5.78 4.58 N/A 6.93 0.63

Sr 475 N/A 436 N/A 443 N/A 466 642 337 477 N/A 434 382 N/A 631 9

Y 12.58 N/A 11.99 N/A 12.30 N/A 12.56 12.08 11.77 13.52 N/A 13.71 11.24 N/A 17.19 1.11

Zr 51.88 N/A 48.59 N/A 54.42 N/A 50.32 54.97 47.73 58.74 N/A 53.78 48.40 N/A 71.87 4.25

Nb 1.98 N/A 1.82 N/A 2.19 N/A 2.26 2.49 1.78 2.40 N/A 1.97 2.07 N/A 3.42 0.45

Ba 147 N/A 133 N/A 147 N/A 160 264 110 153 N/A 138 123 N/A 223 4

La 4.96 N/A 4.65 N/A 5.52 N/A 4.95 7.93 4.00 5.83 N/A 5.03 4.57 N/A 7.55 0.37

Ce 12.07 N/A 11.48 N/A 13.39 N/A 13.84 17.65 10.58 14.32 N/A 13.61 11.64 N/A 17.82 0.85

Pr 1.68 N/A 1.51 N/A 1.84 N/A 1.72 2.28 1.45 1.96 N/A 1.72 1.52 N/A 2.40 0.35

Nd 7.94 N/A 8.05 N/A 8.56 N/A 8.16 10.76 7.22 9.69 N/A 8.15 7.43 N/A 11.35 0.51

Sm 2.21 N/A 2.58 N/A 2.26 N/A 2.01 2.27 1.85 2.63 N/A 2.11 1.94 N/A 2.67 0.37

Eu 0.85 N/A 0.76 N/A 0.89 N/A 0.89 1.00 0.73 0.93 N/A 0.82 0.74 N/A 1.16 0.41

Gd 2.51 N/A 1.96 N/A 2.40 N/A 2.23 2.45 2.06 2.73 N/A 3.13 1.86 N/A 3.07 0.37

Dy 2.18 N/A 2.40 N/A 2.43 N/A 2.50 2.75 2.22 2.84 N/A 2.44 2.00 N/A 3.38 0.42

Er 1.42 N/A 1.39 N/A 1.33 N/A 1.35 1.41 1.31 1.70 N/A 1.27 1.18 N/A 2.01 0.65

Yb 1.49 N/A 1.27 N/A 1.33 N/A 1.30 1.45 1.27 1.57 N/A 1.36 1.07 N/A 1.90 0.40

Hf 1.22 N/A 1.27 N/A 1.37 N/A 1.34 1.51 1.26 1.40 N/A 1.55 1.25 N/A 1.80 3.86

Ta 0.10 N/A 0.11 N/A 0.11 N/A 0.11 0.14 0.10 0.14 N/A 0.13 0.10 N/A 0.16 12.69

Pb 2.29 N/A 2.06 N/A 2.18 N/A 3.20 4.30 1.51 2.44 N/A 2.12 1.83 N/A 2.98 1.34

Th 0.60 N/A 0.58 N/A 0.55 N/A 0.60 1.28 0.49 0.62 N/A 0.62 0.54 N/A 1.01 1.01

U 0.24 N/A 0.26 N/A 0.51 N/A 0.31 0.55 0.16 0.30 N/A 0.27 0.23 N/A 0.37 0.04

Page 149: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

133

BAS-44 B1 B2 B3 B6 B6-2 B7 B9 B17 B18 B20 B22 B23 B24 B25 BB2 BB3 BB4 s.d.

wt%

SiO2 48.89 47.78 48.23 49.50 48.91 49.04 48.51 48.61 49.08 49.01 48.68 49.12 48.78 49.64 48.59 49.15 49.04 0.35

TiO2 0.95 0.93 0.96 0.95 0.96 0.93 0.95 0.95 0.93 0.95 0.93 0.93 0.97 0.95 0.97 0.96 0.97 0.02

Al2O3 16.66 17.81 16.83 15.88 16.50 16.48 16.96 16.43 16.13 16.13 16.50 16.27 16.65 16.37 16.62 16.61 17.04 0.19

FeOT 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 8.88 0.08

MnO 0.13 0.13 0.11 0.13 0.11 0.13 0.11 0.12 0.12 0.12 0.13 0.12 0.11 0.13 0.11 0.12 0.12 0.01

MgO 9.52 9.64 10.13 9.78 9.72 9.86 9.55 10.04 10.21 9.63 10.02 9.82 9.59 10.43 9.76 9.33 9.38 0.07

CaO 10.48 10.48 10.57 10.38 10.53 10.55 10.66 10.44 10.08 10.57 10.45 10.41 10.61 10.66 10.64 10.72 10.34 0.11

Na2O 2.81 2.82 2.72 2.92 2.91 2.54 2.83 2.83 3.02 2.88 2.80 2.76 2.81 1.47 2.80 3.03 2.65 0.05

K2O 0.27 0.29 0.26 0.32 0.28 0.27 0.29 0.30 0.27 0.29 0.27 0.29 0.29 0.27 0.27 0.29 0.34 0.01

P2O5 0.17 0.14 0.16 0.16 0.15 0.16 0.15 0.15 0.16 0.16 0.15 0.16 0.16 0.16 0.15 0.16 0.15 0.01

S 0.11 0.10 0.11 0.11 0.11 0.12 0.12 0.11 0.12 0.12 0.11 0.12 0.12 0.10 0.12 0.11 0.10 0.00

Cl 0.026 0.031 0.029 0.025 0.024 0.026 0.029 0.025 0.029 0.025 0.030 0.025 0.027 0.031 0.033 0.025 0.028 0.00

CO2

(ppm) 494 413 554 446 575 505 654 557 835 615 489 357 784 336 649 90 334 37

H2O 1.14 1.02 1.04 1.01 0.93 1.08 0.99 1.16 1.02 1.26 1.11 1.14 1.07 0.95 1.12 0.64 1.00 0.04

PEC 5.85 7.57 8.19 6.81 6.87 6.91 6.72 6.84 6.21 7.09 6.82 7.43 6.98 7.83 7.05 4.81 6.56

Fo 87.4 87.7 88.2 87.7 87.7 87.7 87.5 88.1 88.3 87.6 88.0 87.7 87.5 87.7 87.8 87.3 87.1

ppm

Li 8.53 N/A 9.37 5.26 5.75 0.63 N/A 6.49 6.23 5.72 7.27 6.21 6.18 N/A 5.84 2.35 N/A 1.49

B 5.75 N/A 2.89 2.55 2.18 2.13 N/A 1.96 2.05 2.04 1.61 3.73 1.72 N/A 2.65 1.67 N/A 1.52

Sc 57.5 N/A 45.2 38.8 39.9 39.0 N/A 43.3 34.9 38.8 38.7 43.2 36.8 N/A 35.8 38.0 N/A 2.9

V 236 N/A 246 211 215 211 N/A 229 180 215 263 222 218 N/A 226 226 N/A 16

Rb 4.47 N/A 7.25 5.19 4.30 4.72 N/A 4.72 4.27 4.54 5.15 5.61 4.59 N/A 4.76 5.93 N/A 1.01

Sr 391 N/A 393 399 408 411 N/A 384 378 413 380 472 404 N/A 396 423 N/A 8

Y 19.59 N/A 19.44 17.88 18.48 19.53 N/A 18.99 17.97 17.72 c 19.35 18.11 N/A 17.99 18.00 N/A 3.00

Zr 79.0 N/A 83.5 77.3 76.6 84.3 N/A 77.6 74.3 72.5 78.9 79.7 74.4 N/A 77.6 79.2 N/A 6.2

Nb 2.57 N/A 2.98 2.67 2.93 2.84 N/A 2.65 2.36 2.59 2.62 3.27 2.63 N/A 2.48 2.79 N/A 1.02

Ba 112 N/A 125 118 111 115 N/A 107 98 108 97 177 109 N/A 104 129 N/A 4

La 7.43 N/A 7.43 7.50 7.57 7.15 N/A 6.96 7.20 7.52 7.26 7.89 7.17 N/A 7.56 7.85 N/A 0.79

Ce 20.6 N/A 22.1 18.6 19.8 18.7 N/A 19.4 17.5 19.4 19.2 21.0 19.2 N/A 20.2 20.5 N/A 1.4

Pr 2.63 N/A 2.83 2.59 2.62 2.72 N/A 2.69 2.58 2.71 2.58 2.83 2.72 N/A 2.61 2.66 N/A 0.83

Nd 12.0 N/A 12.3 11.5 13.0 12.5 N/A 11.9 12.1 12.0 12.2 13.7 11.6 N/A 12.4 12.7 N/A 1.0

Sm 3.0 N/A 2.4 3.0 3.1 3.1 N/A 2.7 2.6 2.9 2.6 3.1 2.8 N/A 3.0 2.9 N/A 1.0

Eu 0.86 N/A 1.09 1.07 1.05 0.97 N/A 1.07 0.94 0.99 1.05 1.12 1.05 N/A 1.18 0.90 N/A 0.85

Gd 3.14 N/A 3.23 3.09 3.22 3.29 N/A 3.18 3.14 3.25 3.20 3.22 2.86 N/A 3.17 3.38 N/A 0.88

Dy 3.34 N/A 3.66 3.36 3.45 3.61 N/A 3.80 3.04 3.11 3.33 3.15 3.31 N/A 3.25 0.00 N/A 1.72

Er 2.22 N/A 2.10 2.12 2.16 1.93 N/A 2.28 2.20 2.30 2.18 2.30 2.17 N/A 2.19 2.33 N/A 0.92

Yb 2.38 N/A 2.13 2.01 2.01 1.91 N/A 2.10 2.14 2.14 1.76 2.20 2.18 N/A 2.02 2.03 N/A 0.82

Hf 1.95 N/A 2.14 1.90 2.02 1.99 N/A 1.94 1.86 1.94 1.97 2.28 1.99 N/A 1.98 1.91 N/A 1.93

Ta 0.15 N/A 0.14 0.15 0.15 0.16 N/A 0.17 0.13 0.15 0.16 0.13 0.12 N/A 0.15 0.17 N/A 3.45

Pb 1.46 N/A 2.07 1.65 1.38 1.35 N/A 1.55 1.62 1.42 1.44 2.00 1.40 N/A 1.46 1.74 N/A 11.91

Th 0.54 N/A 0.61 0.60 0.67 0.59 N/A 0.57 0.58 0.53 0.58 0.62 0.56 N/A 0.60 0.66 N/A 1.71

U 0.20 N/A 0.26 0.21 0.22 0.19 N/A 0.23 0.19 0.18 0.25 0.25 0.19 N/A 0.20 0.29 N/A 0.02

Page 150: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

134

B.2. Boron Isotope Compositions

Sample δ11B 2SE Sample δ11B 2SE

LCC-9-S-

06 -12.63 1.30 BRVB-S-11 -4.69 1.20

LCC-9-S-

07 -11.67 1.24 BRVB-S-14 -5.20 1.08

LCC-9-S-

01 -11.49 1.48 BRVB-S-01 -4.62 1.28

LCC-9-S-

09 -8.79 1.03 BRVB-S-09 -5.38 1.37

LCC-2-S-

07 -5.30 1.86 BPB-S-02 -5.84 1.61

LCC-2-S-

03 -7.98 1.36 BPB-S-04 -3.90 1.62

LCC-2-S-

06 -13.09 1.51 BPB-S-07 -5.53 1.48

LCC-2-S-

10 -8.18 1.32 BRM-S-08 -3.60 1.92

BORG-S-

01 -3.50 2.24 BRM-S-09 -3.17 1.61

BORG-S-

04 -4.21 1.75 BRM-S-03 -2.10 1.54

BORG-S-

10 -3.42 1.81 BRM-S-05 -0.65 1.75

BORG-S-9 1.52 1.68 BPPC-S-04 -4.90 1.60

BORG-S-6 -0.82 1.98 BPPC-S-07 -5.40 1.60

BORG-S-

13 -5.43 1.68 BPPC-S-11 -5.28 1.71

BBL-S-11 -2.89 2.35 BPPC-S-01 -4.23 2.67

BBL-S-06 -7.21 4.16 BAS-44-S-

08 -3.22 4.36

BBL-S-08 -2.78 2.57 BAS-44-S-

06 -6.75 1.55

BBL-S-04 -5.10 3.21 BAS-44-S-

01 -2.78 2.71

BBL-S-05 -3.24 1.10 BAS-44-S-

01 -5.38 2.90

Page 151: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

135

B.3. Radiogenic Isotope Compositions

Sample ID BPCB-1 BORG-1 BPB-1 BBL-05 BAS-44-01 BRVB-1

208Pb/204Pb 38.4640093 38.5386438 38.612363 38.6502638 38.56384767 38.6456054

2SE 0.0027 0.00226 0.00382 0.00206 0.00242 0.00218

207Pb/204Pb 15.5897941 15.6012692 15.612799 15.6183439 15.60589745 15.6166351

2SE 0.001028 0.001022 0.001204 0.000878 0.000878 0.000832

206Pb/204Pb 18.8429231 18.8979534 18.954407 18.9642892 18.92165526 18.977124

2SE 0.00118 0.001138 0.00103 0.001004 0.00091 0.000884

143Nd/144Nd 0.512926 0.51286409 0.5128272 0.51285886 0.512948154 0.51283266

2SE 0.00000594 0.00000642 6.46E-06 0.00000582 0.0000056 0.00000644

176Hf/177Hf 0.28305201 0.28305515 0.2830588 0.28305717 0.283093557 0.28303467

2SE 0.00000722 0.00000646 8.86E-06 0.00000626 0.00000622 0.00000532

87Sr/86Sr 0.70339631 0.70381332 0.7038772 0.70393874 0.703528686 0.70398503

2SE 0.000006 0.000008 0.000008 0.000007 0.000008 0.000008

Page 152: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

136

APPENDIX C

CHAPTER IV SUPPLEMENTARY MATERIALS

C.1. PEC Corrected Melt Inclusion Compositions

Data in this table represent individual melt inclusion compositions that have been

corrected for Fe loss and post entrapment crystallization (PEC). Major elements were

restored using Petrolog 3.1.1 (Danyushevsky and Plechov, 2011). H2O, Cl, S, and trace

elements were restored by the same percent as K2O because they are incompatible with

olivine. H2O values in this table are not corrected for H+ diffusive loss (i.e. they are

different from values shown Fig. 4.6). CO2 values were corrected using methods

described in the text modified from Wallace et al. (2015) and Aster et al. (in prep).

Analyzed CO2 concentrations for these inclusions are reported in C.2. FeO and Fe2O3

were calculated using Petrolog 3.1.1 assuming an fO2 of QFM +1. Temperatures are

calculated by Petrolog 3.1.1 and represent temperatures of melt inclusion entrapment.

Major element standard deviations represent an average of those calculated for each melt

inclusion based on three EPMA point analyses. Trace element standard deviations are

also an average from individual inclusions based on the methods of Loewen and Kent

(2012).

Page 153: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

137

LCC 9-01 9-02 9-06 9-07

09-

09 9-04

9-

04(2) 9-10

9-

16(1)

9-

16(2)

9-

18(1) S.D.

wt%

SiO2 50.64 49.76 49.47 50.23 50.28 50.17 50.41 50.29 49.53 49.43 50.40 0.34

TiO2 0.84 0.91 0.83 0.80 0.77 0.82 0.80 0.78 0.80 0.81 0.80 0.03

Al2O3 16.06 17.37 15.94 16.03 15.88 16.23 16.00 16.21 15.88 15.93 15.85 0.27

Fe2O3 1.05 1.01 1.02 1.01 1.02 1.03 1.03 1.03 1.03 1.02 1.04

FeO 6.06 6.10 6.09 6.10 6.09 6.08 6.08 6.08 6.08 6.09 6.07 0.13

MnO 0.09 0.09 0.08 0.09 0.08 0.08 0.08 0.08 0.10 0.09 0.09 0.01

MgO 8.93 8.42 9.47 9.41 9.51 8.86 8.94 9.29 9.26 9.28 9.20 0.75

CaO 9.74 9.10 10.27 9.76 9.37 9.85 9.55 9.75 10.30 10.34 9.81 0.22

Na2O 3.13 3.32 2.78 2.67 2.74 3.06 2.92 2.92 2.87 2.82 2.96 0.16

K2O 0.83 0.81 0.69 0.79 0.79 0.78 0.85 0.79 0.66 0.66 0.76 0.03

P2O5 0.16 0.17 0.16 0.15 0.15 0.15 0.15 0.15 0.15 0.14 0.15 0.01

S 0.08 0.09 0.09 0.09 0.05 0.09 0.08 0.09 0.09 0.09 0.09 0.003

Cl 0.034 0.037 0.037 0.034 0.037 0.037 0.034 0.036 0.035 0.036 0.035 0.003

H2O 2.59 3.06 3.35 3.11 3.47 3.02 3.32 2.75 3.51 3.53 3.01 0.25

CO2 (ppm)

RESTORED 2180 1011 3298 1252 2101 1371 1727 2912 3030 3101 2792 91

T C 1254 1247 1260 1260 1267 1251 1254 1260 1255 1255 1258

Fo 89.7 89.2 90.2 90.0 90.1 89.6 89.6 90.0 90.0 90.0 89.9

% PEC 9.7 5.0 11.8 4.4 11.5 9.6 10.4 11.2 9.7 10.0 10.2

ppm

Li

5.84 9.69 15.76 7.24 7.60

4.84 7.55

1.93

B

7.20 5.43 6.54 2.30 5.04

3.83 5.32

2.83

Sc

29.83 30.58 31.34 27.11 26.57

29.25 28.73

1.58

V

196 206 178 175 190

192 209

12

Rb

10.20 17.03 14.00 13.74 13.32

9.42 11.75

0.69

Sr

347 371 368 349 339

321 331

9

Y

15.37 14.13 15.67 13.28 12.86

13.25 12.37

0.81

Zr

80.96 75.83 89.94 72.31 67.69

67.79 63.42

4.59

Nb

4.46 4.34 4.06 4.07 3.90

4.17 4.10

0.35

Ba

178 206 207 200 186

157 160

6

La

8.38 7.83 8.08 7.59 7.03

6.57 6.44

0.28

Ce

17.96 17.62 18.47 16.93 16.53

15.46 16.09

0.68

Pr

2.50 2.32 2.74 2.55 2.25

2.18 1.99

0.14

Nd

10.96 10.36 15.07 11.42 10.49

9.21 9.66

0.57

Sm

2.22 2.61 2.07 2.48 2.27

2.26 2.05

0.24

Eu

1.03 0.90 1.09 0.80 0.80

0.79 0.81

0.08

Gd

3.04 2.74 2.87 2.77 2.51

2.30 2.65

0.17

Dy

2.63 2.56 2.66 3.05 2.27

2.45 2.27

0.17

Er

1.66 1.38 2.46 1.22 1.39

1.61 1.52

0.12

Yb

1.43 1.60 1.74 1.56 1.27

1.30 1.48

0.12

Hf

1.77 1.92 3.46 1.92 1.56

1.70 1.15

0.18

Ta

0.28 0.25 0.43 0.18 0.24

0.27 0.26

0.05

Pb

2.40 2.91 2.87 2.90 2.80

1.66 2.13

0.20

Th

1.29 1.65 1.66 1.66 1.53

1.12 0.99

0.12

U 0.35 0.51 0.54 0.58 0.46 0.37 0.37 0.05

Page 154: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

138

LCC-7-

02

LCC-7-

03

LCC-7-

04

LCC-7-

05

LCC-7-

08

LCC-7-

10 S.D.

wt%

SiO2 50.23 50.78 50.48 50.75 50.24 50.27 0.23

TiO2 0.76 0.76 0.78 0.77 0.77 0.78 0.02

Al2O3 16.05 15.86 15.84 15.49 15.86 15.68 0.13

Fe2O3 0.99 1.04 1.04 1.02 1.02 1.03

FeO 6.11 6.07 6.07 6.08 6.08 6.08 0.06

MnO 0.07 0.08 0.08 0.09 0.08 0.08 0.01

MgO 9.73 9.50 9.77 9.70 9.03 9.25 0.12

CaO 10.47 10.22 10.63 9.41 10.06 10.18 0.07

Na2O 2.59 2.95 2.89 2.60 2.92 2.84 0.12

K2O 0.60 0.65 0.61 0.77 0.69 0.66 0.02

P2O5 0.12 0.14 0.13 0.12 0.14 0.14 0.01

S 0.09 0.09 0.08 0.08 0.08 0.07 0.001

Cl 0.038 0.039 0.039 0.037 0.039 0.039 0.001

H2O 2.36 2.04 1.76 3.34 3.24 3.16 0.16

CO2 (ppm)

RESTORED 2801 2356 N/A 1592 2009 2465 75

T C 1260 1261 1264 1269 1252 1255

Fo 90.3 90.2 90.5 90.2 89.7 89.9

% PEC 13.2 10.3 12.0 10.1 9.5 10.5

ppm

Li

17.02 7.11 7.32 1.72

B

4.42 6.37 3.52 2.14

Sc

28.86 29.44 26.14 1.57

V

190 206 220 13

Rb

14.06 13.40 13.18 0.65

Sr

321 338 333 6

Y

13.54 13.84 12.20 0.85

Zr

70.47 72.06 65.83 4.23

Nb

4.02 4.25 4.33 0.34

Ba

195 185 177 5

La

7.39 7.07 6.34 0.23

Ce

15.97 16.51 15.54 0.47

Pr

2.10 2.19 2.04 0.08

Nd

9.08 9.22 9.12 0.39

Sm

2.30 2.49 2.02 0.13

Eu

0.82 0.85 0.75 0.06

Gd

2.53 2.30 2.15 0.22

Dy

2.32 2.63 2.30 0.09

Er

1.35 1.48 1.20 0.09

Yb

1.57 1.38 1.21 0.07

Hf

1.69 1.67 1.65 0.13

Ta

0.19 0.25 0.27 0.03

Pb

3.95 2.82 2.42 0.19

Th

1.27 1.18 1.01 0.09

U 0.43 0.43 0.38 0.03

Page 155: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

139

LCC-6-

01

LCC-6-

02

LCC-6-

06

LCC-6-

07

LCC-6-

08

LCC-6-

10 S.D.

wt%

SiO2 49.92 51.00 50.05 50.57 50.29 53.65 0.31

TiO2 0.82 0.81 0.85 0.84 0.82 0.70 0.02

Al2O3 16.14 16.08 16.38 16.02 16.94 15.67 0.11

Fe2O3 1.03 1.02 1.00 1.02 0.99 1.12

FeO 6.08 6.09 6.10 6.09 6.11 6.00 0.04

MnO 0.07 0.08 0.11 0.10 0.08 0.09 0.01

MgO 9.57 8.83 9.16 8.87 9.02 9.02 0.07

CaO 9.87 9.48 9.70 9.61 9.89 7.24 0.07

Na2O 2.90 2.95 2.82 2.94 2.91 3.15 0.25

K2O 0.70 0.78 0.75 0.78 0.77 1.45 0.02

P2O5 0.15 0.15 0.16 0.16 0.15 0.15 0.01

S 0.09 0.09 0.08 0.09 0.09 0.06 0.003

Cl 0.040 0.037 0.038 0.038 0.039 0.032 0.002

H2O 2.87 2.85 3.04 3.13 2.11 1.84 0.16

CO2 (ppm)

RESTORED 2924 1782 2300 3025 1534 0 77

T C 1266 1251 1255 1251 1250 1284

Fo 90.3 89.4 89.8 89.5 89.7 89.7

% PEC 12.1 11.5 10.1 10.1 10.9 13.1

ppm

Li 9.04 7.04 9.05

35.27 2.62

B 3.56 3.30 5.27

13.09 3.03

Sc 24.70 24.93 26.13

20.54 1.28

V 242 189 196

134 14

Rb 14.32 15.52 15.01

27.66 0.83

Sr 344 351 344

316 8

Y 13.42 13.08 13.44

13.12 0.84

Zr 67.77 68.02 70.46

68.45 4.49

Nb 5.03 4.37 4.63

4.69 0.41

Ba 177 195 189

312 5

La 7.05 7.86 7.32

9.51 0.29

Ce 17.91 18.16 16.88

19.91 0.54

Pr 2.11 2.08 2.07

2.47 0.10

Nd 10.04 9.53 10.09

11.15 0.33

Sm 2.38 2.39 2.34

2.44 0.20

Eu 0.81 0.79 0.81

1.06 0.08

Gd 3.02 2.65 2.68

2.44 0.19

Dy 2.30 2.38 2.15

2.53 0.13

Er 1.64 1.19 1.38

1.36 0.12

Yb 1.48 1.46 1.32

1.73 0.14

Hf 1.80 1.84 1.84

1.72 0.14

Ta 0.28 0.27 0.26

0.34 0.04

Pb 2.88 3.02 2.80

8.29 0.28

Th 1.08 1.58 1.44

3.26 0.14

U 0.38 0.42 0.48 1.62 0.06

Page 156: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

140

LCC

5-

01(MI_1)

5-

01(MI_2) 5-03 5-04 5-05 5-07

5-

11(MI_1)

5-

11(MI_3) 5-12 S.D.

wt%

SiO2 50.83 50.94 52.20 54.25 51.45 51.84 53.89 55.30 51.46 0.22

TiO2 0.81 0.81 0.87 0.77 0.89 0.85 0.67 0.67 0.78 0.02

Al2O3 16.13 15.76 15.90 15.86 15.87 16.29 15.11 15.28 16.57 0.14

Fe2O3 1.05 1.06 1.10 1.13 1.10 1.04 1.23 1.02 1.06

FeO 6.06 6.04 6.01 5.99 6.02 6.06 5.90 6.09 6.05 0.08

MnO 0.09 0.09 0.08 0.10 0.07 0.08 0.10 0.11 0.10 0.02

MgO 8.56 8.56 8.38 8.37 8.61 8.60 8.30 9.28 7.87 0.18

CaO 9.81 9.72 9.60 7.52 9.73 10.04 6.92 6.97 9.56 0.14

Na2O 3.20 3.25 3.55 3.43 3.54 3.21 3.69 1.94 3.41 0.19

K2O 0.85 0.82 0.91 1.46 0.82 0.85 1.77 1.72 1.03 0.02

P2O5 0.16 0.16 0.16 0.16 0.17 0.16 0.13 0.13 0.17 0.01

S 0.09 0.09 0.06 0.08 0.10 0.08 0.06 0.06 0.08 0.003

Cl 0.037 0.036 0.039 0.035 0.036 0.040 0.035 0.035 0.037 0.001

H2O 2.53 2.90 1.28 1.00 1.79 1.03 2.39 1.56 2.01 0.28

CO2 (ppm)

RESTORED 2555 2677 1751 1052 908 2705 0 0 2055 106

T C 1245 1247 1246 1269 1251 1242 1284 1277 1231

Fo 89.3 89.3 89.2 89.1 89.5 89.3 89.3 89.3 88.5

% PEC 8.8 9.6 6.8 12.1 11.4 9.9 11.7 13.8 8.3

ppm

Li

9.17 33.23

51.64

11.13 4.95

B

5.36 10.20

12.39

5.33 4.29

Sc

26.06 18.91

18.28

21.07 1.23

V

186 142

139

180 11

Rb

15.63 28.26

33.25

22.00 1.03

Sr

350 328

296

376 8

Y

13.71 14.07

12.81

12.51 0.83

Zr

76.08 76.37

67.65

70.25 4.45

Nb

4.75 5.21

4.64

4.67 0.42

Ba

208 332

323

243 7

La

8.41 10.56

9.30

8.75 0.27

Ce

18.32 21.67

20.47

19.14 0.56

Pr

2.35 2.52

2.33

2.49 0.09

Nd

9.91 11.45

10.19

10.35 0.40

Sm

2.19 2.41

2.19

2.74 0.20

Eu

0.76 0.77

0.65

0.78 0.06

Gd

2.26 2.55

2.63

2.26 0.17

Dy

2.17 2.49

2.02

2.31 0.13

Er

1.49 1.47

1.39

1.41 0.07

Yb

1.27 1.46

1.74

1.30 0.07

Hf

1.90 1.52

1.72

1.79 0.13

Ta

0.28 0.38

0.31

0.24 0.05

Pb

3.14 9.74

9.82

3.81 0.47

Th

1.60 2.69

2.84

1.46 0.17

U 0.54 1.51 1.63 0.45 0.10

Page 157: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

141

LCC-4-

05

LCC-4-

06(MI_1) LCC-4-13 avg 1 s.d.

wt%

SiO2 52.66 52.08 50.73 0.42

TiO2 0.93 0.94 0.92 0.02

Al2O3 16.64 17.25 16.28 0.07

Fe2O3 1.10 1.03 1.07

FeO 6.02 6.10 6.05 0.04

MnO 0.10 0.08 0.07 0.01

MgO 8.56 7.57 8.65 0.06

CaO 8.33 9.60 9.54 0.09

Na2O 3.25 3.45 3.38 0.06

K2O 1.56 1.00 0.84 0.01

P2O5 0.19 0.18 0.18 0.003

S 0.02 0.06 0.10 0.004

Cl 0.042 0.039 0.046 0.002

H2O 0.71 0.73 2.39 0.28

CO2 (ppm)

RESTORED 1711 3702 2655 99

T C 1262 1219 1252

Fo 89.4 88.0 89.5

% PEC 15.6 9.5 10.5

ppm

Li 6.24 68.95

5.96

B 3.35 23.12

4.57

Sc 20.48 27.77

1.45

V 148 194

13

Rb 11.19 53.70

1.08

Sr 278 461

8

Y 9.84 19.25

1.25

Zr 54.70 105.92

6.77

Nb 3.95 7.92

0.49

Ba 160 515

10

La 6.47 14.92

0.31

Ce 15.21 33.70

0.69

Pr 1.85 3.65

0.13

Nd 7.87 15.52

0.46

Sm 1.87 3.05

0.32

Eu 0.65 1.03

0.08

Gd 1.85 3.98

0.27

Dy 1.92 3.82

0.26

Er 1.10 2.23

0.12

Yb 1.13 2.13

0.14

Hf 1.26 3.29

0.21

Ta 0.28 0.57

0.05

Pb 2.62 14.00

0.57

Th 1.24 4.78

0.17

U 0.32 2.89 0.09

Page 158: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

142

LCC-2-

01

LCC-2-

02

LCC-2-

04

LCC-2-

05

LCC-2-

07

LCC-2-

08

LCC-2-

09

LCC-2-

10 S.D.

wt%

SiO2 50.98 50.96 50.60 50.55 50.57 50.76 50.83 49.99 0.26

TiO2 0.96 0.94 0.95 0.97 0.95 0.94 0.94 0.92 0.02

Al2O3 16.54 16.66 16.57 17.04 16.85 17.03 16.87 16.74 0.11

Fe2O3 1.07 1.00 1.03 1.04 1.06 1.03 1.05 1.05

FeO 6.04 6.10 6.08 6.06 6.05 6.08 6.06 6.06 0.07

MnO 0.09 0.08 0.08 0.09 0.08 0.09 0.09 0.09 0.01

MgO 7.94 8.32 8.24 8.20 8.18 7.66 7.98 8.12 0.05

CaO 9.53 9.07 9.24 9.39 9.47 9.50 9.47 9.41 0.05

Na2O 3.58 2.91 3.22 3.42 3.49 3.43 3.46 3.40 0.14

K2O 0.89 0.95 0.90 0.92 0.91 0.91 0.91 0.89 0.01

P2O5 0.19 0.20 0.20 0.20 0.20 0.20 0.20 0.19 0.01

S 0.09 0.09 0.09 0.09 0.09 0.10 0.09 0.09 0.003

Cl 0.037 0.037 0.038 0.038 0.039 0.039 0.038 0.039 0.002

H2O 2.27 2.93 2.99 2.19 2.27 2.46 2.22 3.25 0.18

CO2 (ppm)

RESTORED 1541 1902 2904 2096 2021 1852 1522 2288 62

T C 1235 1240 1241 1240 1240 1224 1234 1240

Fo 88.7 88.8 88.9 89.0 89.0 88.2 88.7 88.9

% PEC 7.9 9.0 9.2 8.1 8.1 6.2 7.5 8.4

ppm

Li 10.53

7.71

9.59 8.50

1.43

B 6.57

4.77

6.69 6.08

2.70

Sc 26.02

24.16

27.18 27.57

1.37

V 256

191

195 200

13

Rb 24.05

16.82

17.06 17.67

0.78

Sr 392

379

377 388

8

Y 13.90

14.32

14.96 15.32

0.86

Zr 81.28

85.86

86.72 86.62

5.04

Nb 6.69

6.07

6.69 6.31

0.51

Ba 234

228

232 233

6

La 9.13

8.99

9.36 9.53

0.25

Ce 22.40

20.42

21.42 21.04

0.51

Pr 2.66

2.60

2.64 2.69

0.10

Nd 12.04

11.70

12.07 12.03

0.38

Sm 2.54

2.17

2.75 2.86

0.21

Eu 0.84

0.89

0.88 0.90

0.06

Gd 2.66

2.65

2.84 2.61

0.15

Dy 2.77

2.68

2.72 2.59

0.13

Er 1.46

1.73

1.52 1.65

0.13

Yb 1.04

1.70

1.52 1.14

0.16

Hf 1.94

1.72

1.84 2.14

0.15

Ta 0.34

0.33

0.35 0.35

0.04

Pb 3.91

2.95

2.97 3.48

0.19

Th 1.70

1.43

1.61 1.72

0.11

U 0.61 0.48 0.48 0.50 0.04

Page 159: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

143

C.2. Analyzed (uncorrected) Melt Inclusion Compositions

LCC 9-01 9-02 9-06 9-07

9-

09(1)

9-

09(2) 9-04

9-

04(2) 9-10 9-13

9-

16(1)

9-

16(2)

9-

18(1)

1

S.D.

wt%

SiO2 51.04 49.89 50.48 50.57 51.04 51.44 50.89 51.28 51.15 53.14 49.96 50.10 51.18 0.34

Al2O

3 17.42 18.12 17.82 16.72 17.20 17.73 17.72 17.61 17.95 17.70 17.27 17.45 17.39 0.27

FeOT 5.34 5.39 4.99 5.18 4.94 5.05 5.30 5.25 5.18 6.21 5.46 5.31 5.21 0.13

MgO 5.71 7.15 5.72 8.59 5.80 5.88 5.77 5.58 5.67 4.87 6.03 6.05 5.94 0.75

CaO 10.57 9.49 11.48 10.17 10.73 10.46 10.76 10.51 10.79 8.92 11.20 11.32 10.77 0.22

Na2

O 3.39 3.46 3.11 2.78 3.19 3.06 3.34 3.22 3.23 3.72 3.12 3.09 3.24 0.16

K2O 0.90 0.85 0.77 0.82 0.83 0.88 0.85 0.93 0.87 1.35 0.71 0.72 0.83 0.03

TiO2 0.91 0.95 0.93 0.83 0.85 0.86 0.89 0.88 0.86 0.85 0.87 0.89 0.87 0.03

Mn

O 0.10 0.09 0.09 0.09 0.09 0.09 0.09 0.09 0.09 0.11 0.10 0.10 0.10 0.01

P2O5 0.17 0.18 0.18 0.16 0.16 0.16 0.17 0.16 0.17 0.16 0.17 0.16 0.16 0.01

S 0.09 0.09 0.10 0.09 0.09 0.06 0.10 0.09 0.10 0.08 0.10 0.10 0.09 0.003

Cl 0.04 0.04 0.04 0.03 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.003

CO2

(ppm) 965 947 1402 1252 0 848 560 679 1227 648 1500 1545 1339 91

H2O 2.70 3.07 3.59 3.11 3.00 3.71 3.16 3.51 2.92 2.43 3.66 3.71 3.17 0.25

ppm

Li

6.28 9.78 17.32 7.87 8.35

26.82 5.26 8.21

1.93

B

7.74 5.49 7.19 2.50 5.54

10.25 4.17 5.78

2.83

Sc

32.07 30.89 34.44 29.46 29.19

25.71 31.79 31.23

1.58

V

211 209 195 190 209

177 209 227

11.89

Ni

68.29

274.7

5

175.4

0 62.74

232.5

1

47.50 84.43 57.81

20.22

Rb

10.97 17.20 15.39 14.93 14.64

28.57 10.23 12.78

0.69

Sr

373 375 405 379 373

380 349 360

8.53

Y

16.53 14.27 17.22 14.43 14.13

13.86 14.40 13.45

0.81

Zr

87.06 76.60 98.84 78.59 74.38

79.53 73.69 68.94

4.59

Nb

4.80 4.39 4.46 4.42 4.29

4.25 4.53 4.46

0.35

Ba

191 208 227 217 205

305 171 174

5.96

La

9.01 7.90 8.87 8.25 7.73

9.55 7.14 7.00

0.28

Ce

19.32 17.80 20.30 18.40 18.16

20.52 16.80 17.49

0.68

Pr

2.68 2.34 3.02 2.77 2.48

2.53 2.36 2.17

0.14

Nd

11.78 10.46 16.56 12.41 11.53

12.56 10.01 10.50

0.57

Sm

2.38 2.64 2.27 2.70 2.49

2.52 2.46 2.23

0.24

Eu

1.11 0.91 1.19 0.87 0.88

0.86 0.86 0.88

0.08

Gd

3.27 2.77 3.16 3.01 2.76

2.53 2.51 2.88

0.17

Dy

2.82 2.59 2.92 3.32 2.50

2.56 2.66 2.47

0.17

Er

1.79 1.39 2.71 1.33 1.52

1.55 1.75 1.65

0.12

Yb

1.53 1.61 1.92 1.69 1.39

1.43 1.42 1.61

0.12

Hf

1.91 1.94 3.80 2.09 1.71

1.98 1.85 1.26

0.18

Ta

0.30 0.25 0.48 0.20 0.26

0.26 0.29 0.29

0.05

Pb

2.58 2.94 3.15 3.15 3.08

7.18 1.80 2.31

0.20

Th

1.39 1.66 1.82 1.80 1.68

2.48 1.22 1.08

0.12

U 0.38 0.51 0.59 0.63 0.50 0.83 0.40 0.40 0.05

Page 160: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

144

LCC-7-

02

LCC-7-

03

LCC-7-

04

LCC-7-

05

LCC-7-

08

LCC-7-

10

1

S.D.

wt%

SiO2 51.17 51.51 51.47 52.01 51.34 51.48 0.23

Al2O3 18.05 17.38 17.68 17.13 17.43 17.39 0.13

FeOT 4.77 4.96 4.80 5.80 4.89 4.82 0.06

MgO 5.54 6.37 6.07 6.28 6.30 6.18 0.12

CaO 11.78 11.21 11.86 10.40 11.06 11.29 0.07

Na2O 2.91 3.23 3.23 2.87 3.21 3.15 0.12

K2O 0.68 0.72 0.68 0.85 0.76 0.73 0.02

TiO2 0.86 0.83 0.87 0.86 0.85 0.86 0.02

MnO 0.08 0.09 0.08 0.10 0.08 0.09 0.01

P2O5 0.14 0.15 0.15 0.14 0.15 0.16 0.01

S 0.10 0.10 0.08 0.09 0.09 0.07 0.001

Cl 0.04 0.04 0.04 0.04 0.04 0.04 0.001

CO2

(ppm) 974 1159 1014 654 1043 1198 75

H2O 2.54 2.14 1.88 3.54 3.42 3.36 0.16

ppm

Li

18.50 7.64 7.96 1.72

B

4.81 6.85 3.82 2.14

Sc

31.37 31.66 28.41 1.57

V

207 221 239 13

Ni

78.33 107.05 105.84 10.79

Rb

15.28 14.41 14.33 0.65

Sr

349 363 362 6

Y

14.72 14.88 13.26 0.85

Zr

76.60 77.49 71.55 4.23

Nb

4.37 4.57 4.70 0.34

Ba

212 199 192 5

La

8.03 7.60 6.89 0.23

Ce

17.36 17.76 16.89 0.47

Pr

2.29 2.36 2.21 0.08

Nd

9.87 9.91 9.91 0.39

Sm

2.50 2.67 2.20 0.13

Eu

0.89 0.91 0.81 0.06

Gd

2.75 2.47 2.33 0.22

Dy

2.53 2.83 2.50 0.09

Er

1.47 1.59 1.30 0.09

Yb

1.71 1.48 1.31 0.07

Hf

1.83 1.80 1.79 0.13

Ta

0.20 0.27 0.29 0.03

Pb

4.29 3.03 2.63 0.19

Th

1.39 1.27 1.10 0.09

U 0.47 0.46 0.41 0.03

Page 161: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

145

LCC-6-01

LCC-6-

02

LCC-6-

06

LCC-6-

07

LCC-6-

08

LCC-6-

10

1

S.D.

wt%

SiO2 50.72 52.26 50.43 51.18 49.98 55.27 0.31

Al2O3 17.98 17.96 17.84 17.51 18.30 17.70 0.11

FeO 4.96 4.81 5.47 5.23 4.69 5.62 0.04

MgO 5.70 5.27 5.73 5.58 5.56 4.31 0.07

CaO 11.00 10.58 10.56 10.50 10.68 8.17 0.07

Na2O 3.23 3.29 3.07 3.21 3.14 3.56 0.25

K2O 0.78 0.88 0.82 0.86 0.83 1.63 0.02

TiO2 0.91 0.90 0.92 0.91 0.89 0.80 0.02

MnO 0.08 0.08 0.12 0.11 0.08 0.11 0.01

P2O5 0.16 0.17 0.17 0.17 0.16 0.17 0.01

S 0.10 0.10 0.09 0.09 0.10 0.07 0.003

Cl 0.04 0.04 0.04 0.04 0.04 0.04 0.002

CO2

(ppm) 1147 633 970 1426 583 1290 77

H2O 3.06 3.06 3.18 3.29 2.19 1.99 0.16

ppm

Li 10.04 7.73 9.94

40.54 2.62

B 3.95 3.63 5.79

15.05 3.03

Sc 27.45 27.40 28.72

23.61 1.28

V 269 208 215

154 14

Ni 389.75 72.98 148.33

18.70 32.28

Rb 15.91 17.05 16.49

31.80 0.83

Sr 382 385 378

364 8

Y 14.91 14.37 14.77

15.09 0.84

Zr 75.30 74.75 77.43

78.67 4.49

Nb 5.59 4.80 5.08

5.40 0.41

Ba 197 214 207

359 5

La 7.84 8.64 8.05

10.93 0.29

Ce 19.91 19.95 18.55

22.89 0.54

Pr 2.35 2.29 2.28

2.84 0.10

Nd 11.15 10.47 11.08

12.82 0.33

Sm 2.64 2.63 2.57

2.80 0.20

Eu 0.90 0.87 0.89

1.22 0.08

Gd 3.36 2.92 2.95

2.80 0.19

Dy 2.55 2.61 2.36

2.91 0.13

Er 1.82 1.30 1.51

1.57 0.12

Yb 1.64 1.60 1.45

1.98 0.14

Hf 2.00 2.03 2.02

1.98 0.14

Ta 0.31 0.29 0.29

0.39 0.04

Pb 3.19 3.32 3.08

9.53 0.28

Th 1.20 1.74 1.58

3.75 0.14

U 0.42 0.47 0.53 1.86 0.06

Page 162: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

146

5-

01(MI_1)

5-

01(MI_2) 5-03 5-04 5-05 5-07

5-

11(MI_1)

5-

11(MI_3) 5-12 5-13

1

S.D.

wt%

reentrant

SiO2 50.35 50.64 52.53 56.06 52.25 51.00 54.70 55.71 51.28 55.66 0.22

Al2O3 17.08 16.84 16.83 17.83 17.54 17.24 16.66 16.90 17.57 17.23 0.14

FeOT 5.41 5.03 4.79 5.95 4.53 5.01 5.78 6.00 5.93 5.81 0.08

MgO 5.55 5.43 6.68 3.87 5.15 5.27 3.92 3.94 4.78 3.69 0.18

CaO 10.39 10.39 10.16 8.46 10.75 10.63 7.63 7.71 10.13 7.91 0.14

Na2O 3.39 3.47 3.76 3.86 3.91 3.40 4.07 2.15 3.62 4.02 0.19

K2O 0.90 0.88 0.97 1.64 0.91 0.90 1.95 1.90 1.09 1.87 0.02

TiO2 0.86 0.86 0.92 0.86 0.98 0.90 0.74 0.74 0.83 0.74 0.02

MnO 0.10 0.10 0.09 0.11 0.08 0.08 0.11 0.12 0.11 0.11 0.02

P2O5 0.17 0.17 0.17 0.18 0.19 0.17 0.14 0.14 0.18 0.16 0.01

S 0.09 0.09 0.06 0.08 0.10 0.08 0.06 0.06 0.08 0.06 0.003

Cl 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.001

CO2

(ppm) 1259 1283 786 198 1093 1300 1484 317 813 141 106

H2O 2.58 2.98 1.31 1.08 1.90 1.04 2.53 1.65 2.05 1.13 0.28

ppm

Li

10.54 37.77

60.05

12.10 43.23 4.95

B

6.16 11.59

14.41

5.79 14.93 4.29

Sc

29.96 21.49

21.26

22.90 18.54 1.23

V

214 161

161

196 151 11

Ni

280.29 27.14

663.83

215.13 29.76 74.65

Rb

17.96 32.12

38.67

23.92 40.48 1.03

Sr

403 373

344

409 346 8

Y

15.76 15.99

14.90

13.60 14.72 0.83

Zr

87.45 86.78

78.67

76.36 81.45 4.45

Nb

5.46 5.92

5.39

5.08 5.80 0.42

Ba

239 377

375

264 386 7

La

9.67 12.00

10.81

9.51 11.15 0.27

Ce

21.06 24.62

23.80

20.81 23.18 0.56

Pr

2.70 2.87

2.71

2.70 2.83 0.09

Nd

11.39 13.01

11.85

11.25 12.18 0.40

Sm

2.52 2.74

2.55

2.98 2.43 0.20

Eu

0.87 0.88

0.76

0.85 0.80 0.06

Gd

2.60 2.90

3.05

2.46 2.72 0.17

Dy

2.50 2.83

2.35

2.51 2.69 0.13

Er

1.71 1.68

1.61

1.53 1.60 0.07

Yb

1.46 1.66

2.02

1.41 1.73 0.07

Hf

2.18 1.73

2.01

1.95 1.98 0.13

Ta

0.33 0.43

0.36

0.26 0.48 0.05

Pb

3.61 11.07

11.42

4.14 11.59 0.47

Th

1.84 3.05

3.30

1.59 3.71 0.17

U 0.62 1.72 1.89 0.49 2.19 0.10

Page 163: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

147

LCC-4-

01

LCC-4-

05 LCC-4-06(MI_1) LCC-4-06(MI_2)

LCC-4-

08 1 S.D.

wt% reentrant

reentrant

SiO2 52.29 54.49 51.08 55.38 56.67 0.42

Al2O3 18.46 19.23 18.14 17.38 18.22 0.07

FeOT 4.34 5.63 4.79 5.80 5.93 0.04

MgO 3.39 2.75 4.48 3.68 2.86 0.06

CaO 10.67 9.62 10.10 7.77 7.87 0.09

Na2O 4.01 3.76 3.63 4.21 4.15 0.06

K2O 0.91 1.80 1.05 1.94 1.97 0.01

TiO2 1.00 1.07 0.99 0.77 0.81 0.02

MnO 0.08 0.11 0.09 0.11 0.12 0.01

P2O5 0.21 0.22 0.19 0.18 0.18 0.003

S 0.10 0.02 0.06 0.04 0.10 0.004

Cl 0.05 0.05 0.04 0.04 0.04 0.002

CO2

(ppm) 671 251 -105 1323 1046 99

H2O 1.94 0.79 0.74 2.20 2.31 0.28

ppm

Li 63.09 7.52 74.95 11.31

5.96

B 10.38 4.04 25.13 5.43

4.57

Sc 34.09 24.68 30.18 27.77

1.45

V 240 179 211 236

13

Ni 566.96 274.08 2643.00 76.74

138.44

Rb 30.31 13.48 58.37 30.21

1.08

Sr 454 335 501 319

8

Y 19.69 11.86 20.92 15.59

1.25

Zr 109.97 65.91 115.13 73.27

6.77

Nb 7.60 4.76 8.61 4.37

0.49

Ba 394 193 560 578

10

La 12.98 7.80 16.22 9.94

0.31

Ce 29.61 18.33 36.63 24.33

0.69

Pr 3.80 2.23 3.97 3.30

0.13

Nd 16.47 9.48 16.87 14.83

0.46

Sm 4.25 2.25 3.32 3.18

0.32

Eu 1.31 0.79 1.12 0.90

0.08

Gd 3.58 2.23 4.33 3.60

0.27

Dy 3.24 2.31 4.16 3.23

0.26

Er 2.24 1.33 2.43 1.65

0.12

Yb 2.34 1.36 2.31 1.31

0.14

Hf 2.53 1.52 3.57 1.67

0.21

Ta 0.42 0.33 0.62 0.30

0.05

Pb 16.08 3.16 15.22 4.61

0.57

Th 2.50 1.49 5.20 1.68

0.17

U 0.93 0.38 3.14 0.55 0.09

Page 164: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

148

LCC-2-

01

LCC-2-

02

LCC-2-

04

LCC-2-

05 LCC-2-07

LCC-2-

08

LCC-2-

09

LCC-2-

10 1 S.D.

wt%

SiO2 51.36 51.75 51.52 50.85 50.76 50.62 51.20 50.69 0.26

Al2O3 17.69 18.10 18.09 18.24 17.99 17.81 17.99 18.11 0.11

FeOT 4.93 5.12 5.07 5.16 5.35 5.68 5.22 5.22 0.07

MgO 5.68 5.59 5.46 5.71 5.56 5.64 5.75 5.57 0.05

CaO 10.20 9.85 10.09 10.05 10.12 9.93 10.10 10.18 0.05

Na2O 3.83 3.16 3.52 3.66 3.73 3.58 3.69 3.68 0.14

K2O 0.95 1.03 0.98 0.98 0.97 0.95 0.97 0.96 0.01

TiO2 1.02 1.02 1.03 1.04 1.02 0.98 1.01 0.99 0.02

MnO 0.10 0.09 0.09 0.09 0.09 0.09 0.09 0.09 0.01

P2O5 0.20 0.21 0.22 0.22 0.21 0.21 0.21 0.21 0.01

S 0.09 0.10 0.10 0.10 0.10 0.10 0.10 0.10 0.003

Cl 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.002

CO2

(ppm) 819 895 1467 1099 993 1111 812 1168 62

H2O 2.34 3.08 3.16 2.26 2.35 2.49 2.29 3.40 0.18

ppm

Li 11.20

8.38

10.31 9.14

1.43

B 6.99

5.19

7.20 6.54

2.70

Sc 27.68

26.26

29.22 29.65

1.37

V 272

208

210 215

13

Ni 201.94

102.16

1352.67 86.05

36.10

Rb 25.59

18.28

18.34 19.00

0.78

Sr 417

412

406 418

8

Y 14.79

15.57

16.08 16.48

0.86

Zr 86.47

93.33

93.24 93.14

5.04

Nb 7.12

6.60

7.19 6.78

0.51

Ba 249

247

250 250

6

La 9.72

9.77

10.06 10.24

0.25

Ce 23.82

22.19

23.03 22.62

0.51

Pr 2.83

2.83

2.84 2.90

0.10

Nd 12.81

12.72

12.97 12.93

0.38

Sm 2.70

2.36

2.96 3.08

0.21

Eu 0.89

0.97

0.95 0.97

0.06

Gd 2.83

2.88

3.06 2.80

0.15

Dy 2.94

2.92

2.92 2.78

0.13

Er 1.55

1.88

1.64 1.78

0.13

Yb 1.10

1.85

1.63 1.22

0.16

Hf 2.07

1.87

1.98 2.30

0.15

Ta 0.36

0.36

0.38 0.38

0.04

Pb 4.16

3.20

3.19 3.74

0.19

Th 1.81

1.55

1.73 1.85

0.11

U 0.65 0.52 0.52 0.54 0.04

Page 165: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

149

REFERENCES CITED

Chapter I

Anderson, A. T. (1974). Evidence for a picritic, volatile-rich magma beneath Mt. Shasta,

California. Journal of Petrology, 15(2), 243-267.

Cai, Y., LaGatta, A., Goldstein, S. L., Langmuir, C. H., Gómez-Tuena, A., Martín-del

Pozzo, A. L., & Carrasco-Núñez, G. (2014). Hafnium isotope evidence for slab

melt contributions in the Central Mexican Volcanic Belt and implications for slab

melting in hot and cold slab arcs. Chemical Geology, 377, 45-55.

Grove, T., Parman, S., Bowring, S., Price, R., & Baker, M. (2002). The role of an H2O-

rich fluid component in the generation of primitive basaltic andesites and

andesites from the Mt. Shasta region, N. California. Contributions to Mineralogy

and Petrology, 142(4), 375-396.

Hacker, B. R. (2008). H2O subduction beyond arcs. Geochemistry, Geophysics,

Geosystems, 9(3).

Holloway, J. R., & Blank, J. G. (1994). Application of experimental results to COH

species in natural melts. Reviews in Mineralogy and Geochemistry, 30(1), 187-

230.

Johnson, E. R., Wallace, P. J., Cashman, K. V., Granados, H. D., & Kent, A. J. (2008).

Magmatic volatile contents and degassing-induced crystallization at Volcán

Jorullo, Mexico: implications for melt evolution and the plumbing systems of

monogenetic volcanoes. Earth and Planetary Science Letters, 269(3), 478-487.

Jung, H., & Karato, S. I. (2001). Water-induced fabric transitions in

olivine.Science, 293(5534), 1460-1463.

Kessel, R., Schmidt, M. W., Ulmer, P., & Pettke, T. (2005). Trace element signature of

subduction-zone fluids, melts and supercritical liquids at 120–180 km depth.

Nature, 437(7059), 724-727.

Kirby, S., Wang, K., & Dunlop, S. (Eds.). (2002). The Cascadia Subduction Zone and

Related Subduction Systems: Seismic Structure, Intraslab Earthquakes and

Processes, and Earthquake Hazards (pp. 79-80). US Geological Survey.

Leeman, W. P., Smith, D. R., Hildreth, W., Palacz, Z., & Rogers, N. (1990).

Compositional diversity of late Cenozoic basalts in a transect across the southern

Washington Cascades: implications for subduction zone magmatism.Journal of

Geophysical Research: Solid Earth (1978–2012), 95(B12), 19561-19582.

Page 166: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

150

Métrich, N., & Wallace, P. J. (2008). Volatile abundances in basaltic magmas and their

degassing paths tracked by melt inclusions. Reviews in Mineralogy and

Geochemistry, 69(1), 363-402.

Pioli, L., Erlund, E., Johnson, E., Cashman, K., Wallace, P., Rosi, M., & Granados, H. D.

(2008). Explosive dynamics of violent Strombolian eruptions: the eruption of

Parícutin Volcano 1943–1952 (Mexico). Earth and Planetary Science

Letters, 271(1), 359-368.

Plank, T., & Langmuir, C. H. (1993). Tracing trace elements from sediment input to

volcanic output at subduction zones. Nature, 362(6422), 739-743.

Ringwood, A. E. (1974). The petrological evolution of island arc systems Twenty-

seventh William Smith Lecture. Journal of the Geological Society,130(3), 183-

204.

Roggensack, K., Hervig, R. L., McKnight, S. B., & Williams, S. N. (1997). Explosive

basaltic volcanism from Cerro Negro volcano: influence of volatiles on eruptive

style. Science, 277(5332), 1639-1642.

Ruscitto, D. M., Wallace, P. J., Johnson, E. R., Kent, A. J. R., & Bindeman, I. N. (2010).

Volatile contents of mafic magmas from cinder cones in the Central Oregon High

Cascades: Implications for magma formation and mantle conditions in a hot arc.

Earth and Planetary Science Letters, 298(1), 153-161.

Schmidt, M. W., & Poli, S. (1998). Experimentally based water budgets for dehydrating

slabs and consequences for arc magma generation. Earth and Planetary Science

Letters, 163(1), 361-379.

Schiano, P. (2003). Primitive mantle magmas recorded as silicate melt inclusions in

igneous minerals. Earth-Science Reviews, 63(1), 121-144.

Sheppard, P. R., Ort, M. H., Anderson, K. C., Clynne, M. A., & May, E. M. (2009).

Multiple dendrochronological responses to the eruption of Cinder Cone, Lassen

volcanic National Park, California. Dendrochronologia, 27(3), 213-221.

Syracuse, E. M., van Keken, P. E., & Abers, G. A. (2010). The global range of

subduction zone thermal models. Physics of the Earth and Planetary

Interiors,183(1), 73-90.

van Keken, P. E., Hacker, B. R., Syracuse, E. M., & Abers, G. A. (2011). Subduction

factory: 4. Depth‐dependent flux of H2O from subducting slabs worldwide.

Journal of Geophysical Research: Solid Earth (1978–2012), 116(B1).

Wilson, D. S. (2002). The Juan de Fuca plate and slab: Isochron structure and Cenozoic

plate motions. US Geological Survey.

Page 167: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

151

Chapter II

1. Kirby, S. H., Engdahl, E. R., & Villaseñor, A. Warm-slab subduction as a global

process. The Cascadia Subduction Zone and related subduction systems–Seismic

structure, intraslab earthquakes and processes, and earthquake hazards, USGS Open-File

Report, 328, 79-80 (2002).

2. Wada, I., & Wang, K.. Common depth of slab-mantle decoupling: Reconciling

diversity and uniformity of subduction zones. Geochemistry Geophysics Geosystems, 10,

1-36 (2009).

3. Wilson, D.S., The Juan de Fuca plate and slab — isochron structure and Cenozoic

plate motions. U.S. Geological Survey Open-File Report, 328, 9–12 (2002).

4. Rondenay, S., Abers, G. A., & Van Keken, P. E.. Seismic imaging of subduction zone

metamorphism. Geology, 36, 275-278 (2008).

5. Abers, G. A., et al. Imaging the source region of Cascadia tremor and intermediate-

depth earthquakes. Geology, 37, 1119-1122 (2009).

6. Van Keken, P. E., Hacker, B. R., Syracuse, E. M., & Abers, G. A.. Subduction factory:

4. Depth-dependent flux of H2O from subducting slabs worldwide. Journal of

Geophysical Research, 116, B01401 (2011).

7. Ruscitto, D. M., Wallace, P. J., Johnson, E. R., Kent, A. J. R., & Bindeman, I. N.

Volatile contents of mafic magmas from cinder cones in the Central Oregon High

Cascades: Implications for magma formation and mantle conditions in a hot arc. Earth

and Planetary Science Letters, 298, 153–161 (2010).

8. Ruscitto, D. M., Wallace, P. J., & Kent, A. J. R.. Revisiting the compositions and

volatile contents of olivine-hosted melt inclusions from the Mount Shasta region:

implications for the formation of high-Mg andesites. Contributions to Mineralogy and

Petrology, 162, 109-132 (2011).

9. Le Voyer, M., Rose-Koga, E. F., Shimizu, N., Grove, T. L., & Schiano, P. Two

contrasting H2O-rich components in primary melt inclusions from Mount Shasta. Journal

of Petrology, 51, 1571-1595 (2010).

10. Plank, T., Kelley, K. A., Zimmer, M. M., Hauri, E. H., & Wallace, P. J. Why do

mafic arc magmas contain ~4 wt % water on average? Earth and Planetary Science

Letters, 364, 168–179 (2013).

Page 168: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

152

11. Shaw, A. M., Hauri, E. H., Fischer, T. P., Hilton, D. R., & Kelley, K. A. Hydrogen

isotopes in Mariana arc melt inclusions: Implications for subduction dehydration and the

deep-Earth water cycle. Earth and Planetary Science Letters, 275, 138–145 (2008).

12. Leeman, W. P., Tonarini, S., Chan, L. H., & Borg, L. E. Boron and lithium isotopic

variations in a hot subduction zone—the southern Washington Cascades. Chemical

Geology, 212, 101–124 (2004).

13. Borg, L. E., Clynne, M. A., & Bullen, T. D. The variable role of slab-derived fluids in

the generation of a suite of primitive lavas from the Southernmost Cascades. The

Canadian Mineralogist, 35, 425-452 (1997).

14. Borg, L. E., Blichert-Toft, J., Clynne, M. A. Ancient and modern subduction zone

contributions to the mantle sources of lavas from the Lassen Region of California

inferred from Lu – Hf isotopic systematics. Journal of Petrology, 43, 705–723 (2002).

15. Grove, T., Parman, S., Bowring, S., Price, R., & Baker, M. The role of an H2O-rich

fluid component in the generation of primitive basaltic andesites and andesites from the

Mt. Shasta region, N. California. Contributions to Mineralogy and Petrology, 142, 375–

396 (2002).

16. McDonough, W. F. and Sun, S. S. The composition of the Earth. Chemical Geology,

120, 223-253 (1995).

17. Cooper, L. B., et al., Global variations in H2O/Ce: 1. Slab surface temperatures

beneath volcanic arcs. Geochemistry Geophysics Geosystems, 13, 1-27 (2012).

18. McCrory, P. A., Blair, J. L., Waldhauser, F., & Oppenheimer, D. H. Juan de Fuca

slab geometry and its relation to Wadati‐ Benioff zone seismicity. Journal of

Geophysical Research: Solid Earth, 117, 1-16 (2012).

19. Cozzens, B. D., & Spinelli, G. A. A wider seismogenic zone at Cascadia due to fluid

circulation in subducting oceanic crust. Geology, 40, 899-902 (2012).

20. Wilson, C. R., Spiegelman, M., van Keken, P. E., & Hacker, B. R. Fluid flow in

subduction zones: The role of solid rheology and compaction pressure. Earth and

Planetary Science Letters, 401, 261-274, (2014).

21. Schmidt, M. W., & Poli, S. Experimentally based water budgets for dehydrating slabs

and consequences for arc magma generation. Earth and Planetary Science Letters, 163,

361–379 (1998).

22. Hermann, J. & Spandler, C. J. Sediment melts at sub-arc depths: An experimental

study. Journal of Petrology, 49, 717-740 (2008).

Page 169: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

153

23. Gaetani, G. A., O’Leary, J. A., Shimizu, N., Bucholz, C. E., & Newville, M. Rapid

reequilibration of H2O and oxygen fugacity in olivine-hosted melt inclusions. Geology,

40, 915-918 (2012).

24. Bucholz, C. E., Gaetani, G. A., Behn, M. D., & Shimizu, N.. Post-entrapment

modification of volatiles and oxygen fugacity in olivine-hosted melt inclusions. Earth

and Planetary Science Letters, 374, 145–155 (2013).

25. Newman, S., Epstein, S., & Stolper, E. Water, carbon dioxide, and hydrogen isotopes

in glasses from the ca. 1340 AD eruption of the Mono Craters, California: constraints on

degassing phenomena and initial volatile content. Journal of Volcanology and

Geothermal Research, 35, 75-96 (1988).

26. Poreda, R., Schilling, J. G., & Craig, H. Helium and hydrogen isotopes in ocean-ridge

basalts north and south of Iceland. Earth and Planetary Science Letters, 78, 1-17 (1986).

27. Taran, Y. A., Pokrovsky B. G., & Volynets, O. N.. Hydrogen isotopes in hornblendes

and biotites from Quaternary volcanic rocks of the Kamchatka-Kurile arc. Geochemical

Journal – Japan, 31, 203-222 (1997).

28. Giggenbach, W. F. Isotopic shifts in waters from geothermal and volcanic systems

along convergent plate boundaries and their origin. Earth and Planetary Science Letters,

113, 495-510 (1992).

29. Wada, I., Behn, M. D., & Shaw, A. M. Effects of heterogeneous hydration in the

incoming plate, slab rehydration, and mantle wedge hydration on slab-derived H2O flux

in subduction zones. Earth and Planetary Science Letters, 353, 60-71 (2012).

30. Connolly, J. A. D. The geodynamic equation of state: what and how. Geochemistry,

Geophysics, Geosystems, 10, 1-19 (2009).

31. Wannamaker, P. E. et al. Segmentation of plate coupling, fate of subduction fluids,

and modes of arc magmatism in Cascadia, inferred from magnetotelluric resistivity.

Geochemistry, Geophysics, Geosystems, 15, 4230-4253 (2014).

32. Liu, K., Levander, A., Zhai, Y., Porritt, R. W., & Allen, R. M. Asthenospheric flow

and lithospheric evolution near the Mendocino Triple Junction. Earth and Planetary

Science Letters, 323, 60-71 (2012).

33. Alt, J. C., et al. Recycling of water, carbon, and sulfur during subduction of

serpentinites: A stable isotope study of Cerro del Almirez, Spain. Earth and Planetary

Science Letters, 327, 50-60 (2012).

34. Plank, T. Constraints from thorium/lanthanum on sediment recycling at subduction

zones and the evolution of the continents. Journal of Petrology, 46, 921–944 (2005).

Page 170: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

154

35. Davis, A. S., Clague, D. A., Cousens, B. L., Keaten, R., & Paduan, J. B.

Geochemistry of basalt from the North Gorda segment of the Gorda Ridge: Evolution

toward ultraslow spreading ridge lavas due to decreasing magma supply. Geochemistry,

Geophysics, Geosystems, 9, 1-24 (2008).

36. Jicha, B. R. et al. Isotopic and trace element constraints on the petrogenesis of lavas

from the Mount Adams volcanic field, Washington. Contributions to Mineralogy and

Petrology, 157, 189-207 (2009).

37. Mullen, E. K., & Weis, D. Evidence for trench-parallel mantle flow in the northern

Cascade Arc from basalt geochemistry. Earth and Planetary Science Letters, 414, 100-

107 (2015).

38. Skora, S., & Blundy, J. High-pressure hydrous phase relations of radiolarian clay and

implications for the involvement of subducted sediment in arc magmatism. Journal of

Petrology, 51, 2211-2243 (2010).

39. Till, C. B., Grove, T. L., & Withers, A. C. The beginnings of hydrous mantle wedge

melting. Contributions to Mineralogy and Petrology, 163, 669-688 (2012).

40. Spandler, C., & Pirard, C. Element recycling from subducting slabs to arc crust: a

review. Lithos, 170, 208-223 (2013).

41. Defant, M.J., Drummond, M.S. Derivation of some modern arc magmas by melting

of young subducted lithosphere. Nature, 347, 662–665 (1990).

42. Klimm, K., Blundy, J. D., & Green, T. H. Trace element partitioning and accessory

phase saturation during H2O-saturated melting of basalt with implications for subduction

zone chemical fluxes. Journal of Petrology, 49, 523-553 (2008).

43. Jégo, S., & Dasgupta, R. Fluid-present melting of sulfide-bearing ocean-crust:

Experimental constraints on the transport of sulfur from subducting slab to mantle wedge.

Geochimica et Cosmochimica Acta, 110, 106-134 (2013).

44. Hauri, E. H. et al. Matrix effects in hydrogen isotope analysis of silicate glasses by

SIMS. Chemical Geology, 235, 352-365 (2006).

45. Loewen, M. W., & Kent, A. J. Sources of elemental fractionation and uncertainty

during the analysis of semi-volatile metals in silicate glasses using LA-ICP-MS. Journal

of Analytical Atomic Spectrometry, 27, 1502-1508 (2012).

46. Workman, R. K., & Hart, S. R. Major and trace element composition of the depleted

MORB mantle (DMM). Earth and Planetary Science Letters, 231, 53-72 (2005).

Page 171: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

155

47. Donnelly, K. E., Goldstein, S. L., Langmuir, C. H., & Spiegelman, M. Origin of

enriched ocean ridge basalts and implications for mantle dynamics. Earth and Planetary

Science Letters, 226, 347-366 (2004).

48. Shaw, A. M. Long-term preservation of slab signatures in the mantle inferred from

hydrogen isotopes. Nature Geoscience, 5, 224-228 (2012).

49. Hauri, E. H. SIMS analysis of volatiles in silicate glasses, 2: isotopes and abundances

in Hawaiian melt inclusions. Chemical Geology, 183, 115-141 (2002).

Chapter III

Bach, W., Peucker‐Ehrenbrink, B., Hart, S. R., & Blusztajn, J. S. (2003). Geochemistry

of hydrothermally altered oceanic crust: DSDP/ODP Hole 504B–Implications for

seawater‐crust exchange budgets and Sr‐and Pb‐isotopic evolution of the mantle.

Geochemistry, Geophysics, Geosystems, 4(3).

Bacon, C. R., Bruggman, P. E., Christiansen, R. L., Clynne, M. A., Donnelly-Nolan, J.

M., & Hildreth, W. (1997). Primitive magmas at five Cascades volcanic fields:

melts from hot, heterogeneous sub-arc mantle. Canadian Mineralogist, 35, 397-

424.

Baker, M. B., & Stolper, E. M. (1994). Determining the composition of high-pressure

mantle melts using diamond aggregates. Geochimica et Cosmochimica Acta,

58(13), 2811-2827.

Baker, M. B., Grove, T. L., & Price, R. (1994). Primitive basalts and andesites from the

Mt. Shasta region, N. California: products of varying melt fraction and water

content. Contributions to Mineralogy and Petrology, 118(2), 111-129.

Bebout, G. E., Ryan, J. G., & Leeman, W. P. (1993). B-Be systematics in subduction-

related metamorphic rocks: Characterization of the subducted component.

Geochimica et Cosmochimica Acta, 57(10), 2227-2237.

Berge, P. A., & Stauber, D. A. (1987). Seismic refraction study of upper crustal structure

in the Lassen Peak Area, northern California. Journal of Geophysical Research:

Solid Earth (1978–2012), 92(B10), 10571-10579.

Boschi, C., Dini, A., Früh-Green, G. L., & Kelley, D. S. (2008). Isotopic and element

exchange during serpentinization and metasomatism at the Atlantis Massif (MAR

30 N): insights from B and Sr isotope data. Geochimica et Cosmochimica Acta,

72(7), 1801-1823.

Page 172: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

156

Borg, L. E. (1995). The origin and evolution of magmas from the Lassen Region of the

southernmost Cascades. University of Texas at Austin.

Borg, L. E., Clynne, M. A., & Bullen, T. D. (1997). The variable role of slab-derived

fluids in the generation of a suite of primitive calc-alkaline lavas from the

southernmost Cascades, Calilornia. Canadian Mineralogist, 35, 425-452.

Borg, L. E., Brandon, A. D., Clynne, M. A., & Walker, R. J. (2000). Re–Os isotopic

systematics of primitive lavas from the Lassen region of the Cascade arc,

California. Earth and Planetary Science Letters, 177(3), 301-317.

Borg, L. E., Blichert-Toft, J., & Clynne, M. A. (2002). Ancient and modern subduction

zone contributions to the mantle sources of lavas from the Lassen region of

California inferred from Lu–Hf isotopic systematics. Journal of Petrology, 43(4),

705-723.

Bucholz, C. E., Gaetani, G. A., Behn, M. D., & Shimizu, N. (2013). Post-entrapment

modification of volatiles and oxygen fugacity in olivine-hosted melt inclusions.

Earth and Planetary Science Letters, 374, 145-155.

Cagnioncle, A. M., Parmentier, E. M., & Elkins‐Tanton, L. T. (2007). Effect of solid flow

above a subducting slab on water distribution and melting at convergent plate

boundaries. Journal of Geophysical Research: Solid Earth (1978–2012), 112(B9).

Cai, Y., LaGatta, A., Goldstein, S. L., Langmuir, C. H., Gómez-Tuena, A., Martín-del

Pozzo, A. L., & Carrasco-Núñez, G. (2014). Hafnium isotope evidence for slab

melt contributions in the Central Mexican Volcanic Belt and implications for slab

melting in hot and cold slab arcs. Chemical Geology, 377, 45-55.

Carpentier, M., Weis, D., & Chauvel, C. (2013). Large U loss during weathering of upper

continental crust: the sedimentary record. Chemical Geology, 340, 91-104.

Carpentier, M., Weis, D., & Chauvel, C. (2014). Fractionation of Sr and Hf isotopes by

mineral sorting in Cascadia Basin terrigenous sediments. Chemical Geology, 382,

67-82.

Cecil, M. R., Rotberg, G. L., Ducea, M. N., Saleeby, J. B., & Gehrels, G. E. (2012).

Magmatic growth and batholithic root development in the northern Sierra Nevada,

California. Geosphere, 8(3), 592-606.

Chaussidon, M., & Jambon, A. (1994). Boron content and isotopic composition of

oceanic basalts: geochemical and cosmochemical implications. Earth and

Planetary Science Letters, 121(3), 277-291.

Clynne, M. A. (1993). Geologic studies of the Lassen volcanic center, Cascade Range,

California. University of California, Santa Cruz.

Page 173: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

157

Clynne, M. A., & Muffler, L. J. P. (2010). Geologic map of Lassen Volcanic National

Park and vicinity, California. US Department of the Interior, US Geological

Survey.

Clynne, M. A., & Borg, L. E. (1997). Olivine and chromian spinel in primitive calc-

alkaline and tholeiitic lavas from the southernmost Cascade Range, California: a

reflection of relative fertility of the source. Canadian Mineralogist, 35, 453-472.

Cooper, L. B., Ruscitto, D. M., Plank, T., Wallace, P. J., Syracuse, E. M., & Manning, C.

E. (2012). Global variations in H2O/Ce: 1. Slab surface temperatures beneath

volcanic arcs. Geochemistry, Geophysics, Geosystems, 13(3).

Danyushevsky, L. V., & Plechov, P. (2011). Petrolog3: Integrated software for modeling

crystallization processes. Geochemistry, Geophysics, Geosystems, 12(7).

Davis, A. S., Clague, D. A., Cousens, B. L., Keaten, R., & Paduan, J. B. (2008).

Geochemistry of basalt from the North Gorda segment of the Gorda Ridge:

Evolution toward ultraslow spreading ridge lavas due to decreasing magma

supply. Geochemistry, Geophysics, Geosystems, 9(4).

Defant, M. J., & Drummond, M. S. (1990). Derivation of some modern arc magmas by

melting of young subducted lithosphere. Nature, 347(6294), 662-665.

Eiler, J. M. (2001). Oxygen isotope variations of basaltic lavas and upper mantle rocks.

Reviews in Mineralogy and Geochemistry, 43(1), 319-364.

Eiler, J. M., Carr, M. J., Reagan, M., & Stolper, E. (2005). Oxygen isotope constraints on

the sources of Central American arc lavas. Geochemistry, Geophysics,

Geosystems, 6(7).

Gaetani, G. A., O’Leary, J. A., Shimizu, N., Bucholz, C. E., & Newville, M. (2012).

Rapid reequilibration of H2O and oxygen fugacity in olivine-hosted melt

inclusions. Geology, 40(10), 915-918.

Garth, T., & Rietbrock, A. (2014). Order of magnitude increase in subducted H2O due to

hydrated normal faults within the Wadati-Benioff zone. Geology, 42(3), 207-210.

Ghiorso, M. S., Hirschmann, M. M., Reiners, P. W., & Kress, V. C. (2002). The

pMELTS: A revision of MELTS for improved calculation of phase relations and

major element partitioning related to partial melting of the mantle to 3 GPa.

Geochemistry, Geophysics, Geosystems, 3(5), 1-35.

Page 174: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

158

Grove, T., Parman, S., Bowring, S., Price, R., & Baker, M. (2002). The role of an H2O-

rich fluid component in the generation of primitive basaltic andesites and

andesites from the Mt. Shasta region, N. California. Contributions to Mineralogy

and Petrology, 142(4), 375-396.

Guffanti, M., & Weaver, C. S. (1988). Distribution of late Cenozoic volcanic vents in the

Cascade Range: Volcanic arc segmentation and regional tectonic considerations.

Journal of Geophysical Research: Solid Earth (1978–2012), 93(B6), 6513-6529.

Guffanti, M., Clynne, M. A., Smith, J. G., Muffler, L. J. P., & Bullen, T. D. (1990). Late

Cenozoic volcanism, subduction, and extension in the Lassen region of

California, southern Cascade Range. Journal of Geophysical Research: Solid

Earth (1978–2012), 95(B12), 19453-19464.

Hacker, B. R. (2008). H2O subduction beyond arcs. Geochemistry, Geophysics,

Geosystems, 9(3).

Hermann, J., Spandler, C., Hack, A., & Korsakov, A. V. (2006). Aqueous fluids and

hydrous melts in high-pressure and ultra-high pressure rocks: implications for

element transfer in subduction zones. Lithos, 92(3), 399-417.

Hirschmann, M. M., Baker, M. B., & Stolper, E. M. (1998). The effect of alkalis on the

silica content of mantle-derived melts. Geochimica et Cosmochimica Acta, 62(5),

883-902.

Hughes, S. S., & Taylor, E. M. (1986). Geochemistry, petrogenesis, and tectonic

implications of central High Cascade mafic platform lavas. Geological Society of

America Bulletin, 97(8), 1024-1036.

Ishikawa, T., & Nakamura, E. (1993). Boron isotope systematics of marine sediments.

Earth and Planetary Science Letters, 117(3), 567-580.

Ishikawa, T., & Tera, F. (1999). Two isotopically distinct fluid components involved in

the Mariana arc: Evidence from Nb/B ratios and B, Sr, Nd, and Pb isotope

systematics. Geology, 27(1), 83-86.

Ishikawa, T., Tera, F., & Nakazawa, T. (2001). Boron isotope and trace element

systematics of the three volcanic zones in the Kamchatka arc. Geochimica et

Cosmochimica Acta, 65(24), 4523-4537.

Jégo, S., & Dasgupta, R. (2013). Fluid-present melting of sulfide-bearing ocean-crust:

Experimental constraints on the transport of sulfur from subducting slab to mantle

wedge. Geochimica et Cosmochimica Acta, 110, 106-134.

Page 175: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

159

Johnson, E. R., Wallace, P. J., Cashman, K. V., Granados, H. D., & Kent, A. J. (2008).

Magmatic volatile contents and degassing-induced crystallization at Volcán

Jorullo, Mexico: implications for melt evolution and the plumbing systems of

monogenetic volcanoes. Earth and Planetary Science Letters, 269(3), 478-487.

Kelemen, P. B., Shimizu, N., & Dunn, T. (1993). Relative depletion of niobium in some

arc magmas and the continental crust: partitioning of K, Nb, La and Ce during

melt/rock reaction in the upper mantle. Earth and Planetary Science Letters,

120(3), 111-134.

Kessel, R., Schmidt, M. W., Ulmer, P., & Pettke, T. (2005). Trace element signature of

subduction-zone fluids, melts and supercritical liquids at 120–180 km depth.

Nature, 437(7059), 724-727.

Kirby, S., Wang, K., & Dunlop, S. (Eds.). (2002). The Cascadia Subduction Zone and

Related Subduction Systems: Seismic Structure, Intraslab Earthquakes and

Processes, and Earthquake Hazards (pp. 79-80). US Geological Survey.

Klimm, K., Blundy, J. D., & Green, T. H. (2008). Trace element partitioning and

accessory phase saturation during H2O-saturated melting of basalt with

implications for subduction zone chemical fluxes. Journal of Petrology, 49(3),

523-553.

Krawczynski, M. J., Grove, T. L., & Behrens, H. (2012). Amphibole stability in primitive

arc magmas: effects of temperature, H2O content, and oxygen fugacity.

Contributions to Mineralogy and Petrology, 164(2), 317-339.

Leeman, W. P., Tonarini, S., Chan, L. H., & Borg, L. E. (2004). Boron and lithium

isotopic variations in a hot subduction zone—the southern Washington Cascades.

Chemical Geology, 212(1), 101-124.

Liu, K., Levander, A., Zhai, Y., Porritt, R. W., & Allen, R. M. (2012). Asthenospheric

flow and lithospheric evolution near the Mendocino Triple Junction. Earth and

Planetary Science Letters, 323, 60-71.

Lloyd, A. S., Plank, T., Ruprecht, P., Hauri, E. H., & Rose, W. (2013). Volatile loss from

melt inclusions in pyroclasts of differing sizes. Contributions to Mineralogy and

Petrology, 165(1), 129-153.

Loewen, M. W., & Kent, A. J. (2012). Sources of elemental fractionation and uncertainty

during the analysis of semi-volatile metals in silicate glasses using LA-ICP-MS.

Journal of Analytical Atomic Spectrometry, 27(9), 1502-1508.

Le Voyer, M., Rose-Koga, E. F., Shimizu, N., Grove, T. L., & Schiano, P. (2010). Two

contrasting H2O-rich components in primary melt inclusions from Mount Shasta.

Journal of Petrology, 51(7), 1571-1595.

Page 176: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

160

Manea, V. C., Leeman, W. P., Gerya, T., Manea, M., & Zhu, G. (2014). Subduction of

fracture zones controls mantle melting and geochemical signature above slabs.

Nature communications, 5.

Manning, C. E. (2004). The chemistry of subduction-zone fluids. Earth and Planetary

Science Letters, 223(1), 1-16.

Marschall, H. R., & Monteleone, B. D. (2014). Boron Isotope Analysis of Silicate Glass

with Very Low Boron Concentrations by Secondary Ion Mass Spectrometry.

Geostandards and Geoanalytical Research.

Marschall, H. R., Altherr, R., & Rüpke, L. (2007). Squeezing out the slab—modelling the

release of Li, Be and B during progressive high-pressure metamorphism.

Chemical Geology, 239(3), 323-335.

Mullen, E. K., & Weis, D. (2015). Evidence for trench-parallel mantle flow in the

northern Cascade Arc from basalt geochemistry. Earth and Planetary Science

Letters, 414, 100-107.

Nedimović, M. R., Bohnenstiehl, D. R., Carbotte, S. M., Canales, J. P., & Dziak, R. P.

(2009). Faulting and hydration of the Juan de Fuca plate system. Earth and

Planetary Science Letters, 284(1), 94-102.

Peccerillo, A., & Taylor, S. R. (1976). Geochemistry of Eocene calc-alkaline volcanic

rocks from the Kastamonu area, northern Turkey. Contributions to mineralogy

and petrology, 58(1), 63-81.

Plank, T. (2005). Constraints from thorium/lanthanum on sediment recycling at

subduction zones and the evolution of the continents. Journal of Petrology, 46(5),

921-944.

Plank, T., Cooper, L. B., & Manning, C. E. (2009). Emerging geothermometers for

estimating slab surface temperatures. Nature Geoscience, 2(9), 611-615.

Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., & Khubunaya, S. (2007).

Constraints on mantle melting and composition and nature of slab components in

volcanic arcs from volatiles (H2O, S, Cl, F) and trace elements in melt inclusions

from the Kamchatka Arc. Earth and Planetary Science Letters, 255(1), 53-69.

Ranero, C. R., Morgan, J. P., McIntosh, K., & Reichert, C. (2003). Bending-related

faulting and mantle serpentinization at the Middle America trench. Nature,

425(6956), 367-373.

Page 177: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

161

Rose, E. F., Shimizu, N., Layne, G. D., & Grove, T. L. (2001). Melt production beneath

Mt. Shasta from boron data in primitive melt inclusions. Science, 293(5528), 281-

283.

Ryan, J. G., & Langmuir, C. H. (1993). The systematics of boron abundances in young

volcanic rocks. Geochimica et Cosmochimica Acta, 57(7), 1489-1498.

Ruscitto, D. M., Wallace, P. J., Johnson, E. R., Kent, A. J. R., & Bindeman, I. N. (2010).

Volatile contents of mafic magmas from cinder cones in the Central Oregon High

Cascades: Implications for magma formation and mantle conditions in a hot arc.

Earth and Planetary Science Letters, 298(1), 153-161.

Ruscitto, D. M., Wallace, P. J., & Kent, A. J. R. (2011). Revisiting the compositions and

volatile contents of olivine-hosted melt inclusions from the Mount Shasta region:

implications for the formation of high-Mg andesites. Contributions to Mineralogy

and Petrology, 162(1), 109-132.

Ruscitto, D. M., Wallace, P. J., Cooper, L. B., & Plank, T. (2012). Global variations in

H2O/Ce: 2. Relationships to arc magma geochemistry and volatile fluxes.

Geochemistry, Geophysics, Geosystems, 13(3).

Schmidt, M. W., & Poli, S. (1998). Experimentally based water budgets for dehydrating

slabs and consequences for arc magma generation. Earth and Planetary Science

Letters, 163(1), 361-379.

Sherrod, D. R., & Smith, J. G. (1990). Quaternary extrusion rates of the Cascade Range,

northwestern United States and southern British Columbia. Journal of

Geophysical Research: Solid Earth (1978–2012), 95(B12), 19465-19474.

Spandler, C., & Pirard, C. (2013). Element recycling from subducting slabs to arc crust: a

review. Lithos, 170, 208-223.

Spandler, C., Pettke, T., & Hermann, J. (2014). Experimental study of trace element

release during ultrahigh-pressure serpentinite dehydration. Earth and Planetary

Science Letters, 391, 296-306.

Smith, H. J., Spivack, A. J., Staudigel, H., & Hart, S. R. (1995). The boron isotopic

composition of altered oceanic crust. Chemical Geology, 126(2), 119-135.

Sun, S. S., & McDonough, W. F. (1989). Chemical and isotopic systematics of oceanic

basalts: implications for mantle composition and processes. Geological Society,

London, Special Publications, 42(1), 313-345.

Syracuse, E. M., van Keken, P. E., & Abers, G. A. (2010). The global range of

subduction zone thermal models. Physics of the Earth and Planetary

Interiors,183(1), 73-90.

Page 178: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

162

Till, C. B., Grove, T. L., Carlson, R. W., Donnelly‐Nolan, J. M., Fouch, M. J., Wagner,

L. S., & Hart, W. K. (2013). Depths and temperatures of< 10.5 Ma mantle

melting and the lithosphere‐asthenosphere boundary below southern Oregon and

northern California. Geochemistry, Geophysics, Geosystems, 14(4), 864-879.

Tonarini, S., Armienti, P., D’Orazio, M., & Innocenti, F. (2001). Subduction-like fluids

in the genesis of Mt. Etna magmas: evidence from boron isotopes and fluid

mobile elements. Earth and Planetary Science Letters, 192(4), 471-483.

Wada, I., & Wang, K. (2009). Common depth of slab‐mantle decoupling: Reconciling

diversity and uniformity of subduction zones. Geochemistry, Geophysics,

Geosystems, 10(10).

Wada, I., Behn, M. D., & Shaw, A. M. (2012). Effects of heterogeneous hydration in the

incoming plate, slab rehydration, and mantle wedge hydration on slab-derived

H2O flux in subduction zones. Earth and Planetary Science Letters, 353, 60-71.

Wannamaker, P. E., Booker, J. R., Jones, A. G., Chave, A. D., Filloux, J. H., Waff, H. S.,

& Law, L. K. (1989). Resistivity cross section through the Juan de Fuca

subduction system and its tectonic implications. Journal of Geophysical

Research: Solid Earth (1978–2012), 94(B10), 14127-14144.

Weis, D., Kieffer, B., Maerschalk, C., Barling, J., De Jong, J., Williams, G. A., ... &

Mahoney, J. B. (2006). High‐precision isotopic characterization of USGS

reference materials by TIMS and MC‐ICP‐MS. Geochemistry, Geophysics,

Geosystems, 7(8).

Weis, D., Kieffer, B., Hanano, D., Nobre Silva, I., Barling, J., Pretorius, W. & Mattielli,

N. (2007). Hf isotope compositions of US Geological Survey reference materials.

Geochemistry, Geophysics, Geosystems, 8(6).

Wilson, D. S. (2002). The Juan de Fuca plate and slab: Isochron structure and Cenozoic

plate motions. US Geological Survey Open-File Report, 02-328.

White, S. M., Crisp, J. A., & Spera, F. J. (2006). Long‐term volumetric eruption rates and

magma budgets. Geochemistry, Geophysics, Geosystems, 7(3).

Workman, R. K., & Hart, S. R. (2005). Major and trace element composition of the

depleted MORB mantle (DMM). Earth and Planetary Science Letters, 231(1),

53-72.

van Keken, P. E., Hacker, B. R., Syracuse, E. M., & Abers, G. A. (2011). Subduction

factory: 4. Depth‐dependent flux of H2O from subducting slabs worldwide.

Journal of Geophysical Research: Solid Earth (1978–2012), 116(B1).

Page 179: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

163

Walowski, K. J., Wallace, P. J., Hauri, E. H., Wada, I., & Clynne, M. A. (2015). Slab

melting beneath the Cascade Arc driven by dehydration of altered oceanic

peridotite. Nature Geoscience, 8(5), 404-408.

Yogodzinski, G. M., Kay, R. W., Volynets, O. N., Koloskov, A. V., & Kay, S. M. (1995).

Magnesian andesite in the western Aleutian Komandorsky region: implications

for slab melting and processes in the mantle wedge. Geological Society of

America Bulletin, 107(5), 505-519.

Chapter IV

Anderson, A. T. (1974). Evidence for a picritic, volatile-rich magma beneath Mt. Shasta,

California. Journal of Petrology, 15(2), 243-267.

Beattie, P. (1994). Systematics and energetics of trace-element partitioning between

olivine and silicate melts: implications for the nature of mineral/melt

partitioning. Chemical Geology, 117(1), 57-71.

Berge, P. A., & Stauber, D. A. (1987). Seismic refraction study of upper crustal structure

in the Lassen Peak Area, northern California. Journal of Geophysical Research:

Solid Earth (1978–2012), 92(B10), 10571-10579.

Blatter, D. L., Sisson, T. W., & Hankins, W. B. (2013). Crystallization of oxidized,

moderately hydrous arc basalt at mid-to lower-crustal pressures: implications for

andesite genesis. Contributions to Mineralogy and Petrology,166(3), 861-886.

Brenna, M., Cronin, S. J., Nemeth, K., Smith, I. E., & Sohn, Y. K. (2011). The influence

of magma plumbing complexity on monogenetic eruptions, Jeju Island,

Korea. Terra Nova, 23(2), 70-75.

Brenna, M., Cronin, S. J., Smith, I. E., Sohn, Y. K., & Németh, K. (2010). Mechanisms

driving polymagmatic activity at a monogenetic volcano, Udo, Jeju Island, South

Korea. Contributions to Mineralogy and Petrology, 160(6), 931-950.

Borg, L. E., Clynne, M. A., & Bullen, T. D. (1997). The variable role of slab-derived

fluids in the generation of a suite of primitive calc-alkaline lavas from the

southernmost Cascades, Calilornia. Canadian Mineralogist, 35, 425-452.

Borg, L. E., Blichert-Toft, J., & Clynne, M. A. (2002). Ancient and modern subduction

zone contributions to the mantle sources of lavas from the Lassen region of

California inferred from Lu–Hf isotopic systematics. Journal of Petrology, 43(4),

705-723.

Page 180: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

164

Bucholz, C. E., Gaetani, G. A., Behn, M. D., & Shimizu, N. (2013). Post-entrapment

modification of volatiles and oxygen fugacity in olivine-hosted melt

inclusions. Earth and Planetary Science Letters, 374, 145-155.

Cecil, M. R., Rotberg, G. L., Ducea, M. N., Saleeby, J. B., & Gehrels, G. E. (2012).

Magmatic growth and batholithic root development in the northern Sierra Nevada,

California. Geosphere, 8(3), 592-606.

Clynne, M. A. (1993). Geologic studies of the Lassen volcanic center, Cascade Range,

California. University of California, Santa Cruz.

Clynne, M. A. (2011, December). Stratigraphy and compositional evolution of Cinder

Cone, a composite monogenetic cone in Lassen Volcanic National Park,

California. In AGU Fall Meeting Abstracts (Vol. 1, p. 2621).

Clynne, M. A., & Muffler, L. J. P. (2010). Geologic map of Lassen Volcanic National

Park and vicinity, California. US Department of the Interior, US Geological

Survey.

Danyushevsky, L. V., Della-Pasqua, F. N., & Sokolov, S. (2000). Re-equilibration of

melt inclusions trapped by magnesian olivine phenocrysts from subduction-

related magmas: petrological implications. Contributions to Mineralogy and

Petrology, 138(1), 68-83.

Danyushevsky, L. V., & Plechov, P. (2011). Petrolog3: Integrated software for modeling

crystallization processes. Geochemistry, Geophysics, Geosystems,12(7).

Dixon, J. E., & Pan, V. (1995). Determination of the molar absorptivity of dissolved

carbonate in basanitic glass. American Mineralogist, 80(11-12), 1339-1342.

Dohmen, R., & Chakraborty, S. (2007). Fe–Mg diffusion in olivine II: point defect

chemistry, change of diffusion mechanisms and a model for calculation of

diffusion coefficients in natural olivine. Physics and Chemistry of Minerals,34(6),

409-430.

Erlund, E. J., Cashman, K. V., Wallace, P. J., Pioli, L., Rosi, M., Johnson, E., &

Granados, H. D. (2010). Compositional evolution of magma from Parícutin

Volcano, Mexico: the tephra record. Journal of Volcanology and Geothermal

Research, 197(1), 167-187.

Esposito, R., Bodnar, R. J., Danyushevsky, L. V., De Vivo, B., Fedele, L., Hunter, J., ...

& Shimizu, N. (2011). Volatile evolution of magma associated with the Solchiaro

eruption in the Phlegrean Volcanic District (Italy). Journal of Petrology, 52(12),

2431-2460.

Page 181: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

165

Ford, C. E., Russell, D. G., Craven, J. A., & Fisk, M. R. (1983). Olivine-liquid equilibria:

temperature, pressure and composition dependence of the crystal/liquid cation

partition coefficients for Mg, Fe2+, Ca and Mn. Journal of Petrology, 24(3), 256-

266.

Gaetani, G. A., & Grove, T. L. (1998). The influence of water on melting of mantle

peridotite. Contributions to Mineralogy and Petrology, 131(4), 323-346.

Gaetani, G. A., O’Leary, J. A., Shimizu, N., Bucholz, C. E., & Newville, M. (2012).

Rapid reequilibration of H2O and oxygen fugacity in olivine-hosted melt

inclusions. Geology, 40(10), 915-918.

Ghiorso, M. S., Hirschmann, M. M., Reiners, P. W., & Kress, V. C. (2002). The

pMELTS: A revision of MELTS for improved calculation of phase relations and

major element partitioning related to partial melting of the mantle to 3 GPa.

Geochemistry, Geophysics, Geosystems, 3(5), 1-35.

Grove, T. L., Kinzler, R. J., Baker, M. B., Donnelly-Nolan, J. M., & Lesher, C. E. (1988).

Assimilation of granite by basaltic magma at Burnt Lava flow, Medicine Lake

volcano, northern California: decoupling of heat and mass transfer. Contributions

to Mineralogy and Petrology, 99(3), 320-343.

Grove, T. L., Chatterjee, N., Parman, S. W., & Médard, E. (2006). The influence of H2O

on mantle wedge melting. Earth and Planetary Science Letters, 249(1), 74-89.

Gualda, G. A., Ghiorso, M. S., Lemons, R. V., & Carley, T. L. (2012). Rhyolite-MELTS:

a modified calibration of MELTS optimized for silica-rich, fluid-bearing

magmatic systems. Journal of Petrology, 53(5), 875-890.

Guffanti, M., Clynne, M. A., Smith, J. G., Muffler, L. J. P., & Bullen, T. D. (1990). Late

Cenozoic volcanism, subduction, and extension in the Lassen region of

California, southern Cascade Range. Journal of Geophysical Research: Solid

Earth (1978–2012), 95(B12), 19453-19464.

Hartley, M. E., Maclennan, J., Edmonds, M., & Thordarson, T. (2014). Reconstructing

the deep CO2 degassing behaviour of large basaltic fissure eruptions. Earth and

Planetary Science Letters, 393, 120-131.

Hauri, E. H., Wagner, T. P., & Grove, T. L. (1994). Experimental and natural partitioning

of Th, U, Pb and other trace elements between garnet, clinopyroxene and basaltic

melts. Chemical Geology, 117(1), 149-166.

Heiken, G. (1978). Characteristics of tephra from cinder cone, Lassen volcanic national

park, California. Bulletin Volcanologique, 41(2), 119-130.

Page 182: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

166

Hildreth, W. (2007). Quaternary magmatism in the Cascades-geologic perspectives. US

Geological Survey professional paper, (1744), 1-125.

Hirschmann, M. M., Tenner, T., Aubaud, C., & Withers, A. C. (2009). Dehydration

melting of nominally anhydrous mantle: the primacy of partitioning. Physics of

the Earth and Planetary Interiors, 176(1), 54-68.

Holloway, J. R., & Blank, J. G. (1994). Application of experimental results to COH

species in natural melts. Reviews in Mineralogy and Geochemistry, 30(1), 187-

230.

Iacono-Marziano, G., Morizet, Y., Le Trong, E., & Gaillard, F. (2012). New experimental

data and semi-empirical parameterization of H2O–CO2 solubility in mafic

melts. Geochimica et Cosmochimica Acta, 97, 1-23.

Johnson, E. R., Wallace, P. J., Cashman, K. V., Granados, H. D., & Kent, A. J. (2008).

Magmatic volatile contents and degassing-induced crystallization at Volcán

Jorullo, Mexico: implications for melt evolution and the plumbing systems of

monogenetic volcanoes. Earth and Planetary Science Letters, 269(3), 478-487.

Johnson, E. R., Wallace, P. J., Cashman, K. V., & Granados, H. D. (2010). Degassing of

volatiles (H 2 O, CO 2, S, Cl) during ascent, crystallization, and eruption at mafic

monogenetic volcanoes in central Mexico. Journal of Volcanology and

Geothermal Research, 197(1), 225-238.

Kumazawa, M., & Anderson, O. L. (1969). Elastic moduli, pressure derivatives, and

temperature derivatives of single‐crystal olivine and single‐crystal forsterite.

Journal of Geophysical Research, 74(25), 5961-5972.

Lange, R. L., & Carmichael, I. S. (1990). Thermodynamic properties of silicate liquids

with emphasis on density, thermal expansion and compressibility. Reviews in

Mineralogy and Geochemistry, 24(1), 25-64.

Le Bas, M. J., Le Maitre, R. W., Streckeisen, A., & Zanettin, B. (1986). A chemical

classification of volcanic rocks based on the total alkali-silica diagram. Journal of

Petrology, 27(3), 745-750.

Le Voyer, M., Rose-Koga, E. F., Shimizu, N., Grove, T. L., & Schiano, P. (2010). Two

contrasting H2O-rich components in primary melt inclusions from Mount Shasta.

Journal of Petrology, 51(7), 1571-1595.

Leeman, W. P., Tonarini, S., Chan, L. H., & Borg, L. E. (2004). Boron and lithium

isotopic variations in a hot subduction zone—the southern Washington

Cascades. Chemical Geology, 212(1), 101-124.

Page 183: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

167

Lloyd, A. S., Plank, T., Ruprecht, P., Hauri, E. H., & Rose, W. (2013). Volatile loss from

melt inclusions in pyroclasts of differing sizes. Contributions to Mineralogy and

Petrology, 165(1), 129-153.

Loewen, M. W., & Kent, A. J. (2012). Sources of elemental fractionation and uncertainty

during the analysis of semi-volatile metals in silicate glasses using LA-ICP-

MS. Journal of Analytical Atomic Spectrometry, 27(9), 1502-1508.

Lowenstern, J. B. (1995). Applications of silicate-melt inclusions to the study of

magmatic volatiles. Magmas, Fluids and Ore Deposits. Mineralogical

Association of Canada Short Course, 23, 71-99.

Luhr, J. F. (2001). Glass inclusions and melt volatile contents at Paricutin Volcano,

Mexico. Contributions to Mineralogy and Petrology, 142(3), 261-283.

Luhr, J. F., & Carmichael, I. S. (1985). Jorullo Volcano, Michoacán, Mexico (1759–

1774): the earliest stages of fractionation in calc-alkaline magmas. Contributions

to Mineralogy and Petrology, 90(2-3), 142-161.

McBirney, A. R., Taylor, H. P., & Armstrong, R. L. (1987). Paricutin re-examined: a

classic example of crustal assimilation in calc-alkaline magma.Contributions to

Mineralogy and Petrology, 95(1), 4-20.

Mallmann, G., & O’Neill, H. S. C. (2009). The crystal/melt partitioning of V during

mantle melting as a function of oxygen fugacity compared with some other

elements (Al, P, Ca, Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). Journal of

Petrology, 50(9), 1765-1794.

Marks, J.K. (2012). Physical Volcanology of the 1666 C.E. Cinder Cone Eruption,

Lassen Volcanic National Park, California. University of Oregon.

McKenzie, D. A. N., & O'nions, R. K. (1991). Partial melt distributions from inversion of

rare earth element concentrations. Journal of Petrology, 32(5), 1021-1091.

Métrich, N., & Wallace, P. J. (2008). Volatile abundances in basaltic magmas and their

degassing paths tracked by melt inclusions. Reviews in Mineralogy and

Geochemistry, 69(1), 363-402.

Moore, L. R. Gazel, E., Tuohy, R., Lloyd, A., Esposito, R., Steele-MacInnis, M., Hauri,

E.H., Wallace, P.J., Plank, T., Bodnar, R.J. (2014). Bubbles matter: An

assessment of the contribution of vapor bubbles to melt inclusion volatile budgets.

The American Mineralogist, in press.

Newman, S., & Lowenstern, J. B. (2002). VolatileCalc: a silicate melt–H2O–CO2 solution

model written in Visual Basic for Excel. Computers & Geosciences, 28(5), 597-

604.

Page 184: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

168

Papale, P., Moretti, R., & Barbato, D. (2006). The compositional dependence of the

saturation surface of H2O + CO2 fluids in silicate melts. Chemical

Geology, 229(1), 78-95.

Paster, T. P., Schauwecker, D. S., & Haskin, L. A. (1974). The behavior of some trace

elements during solidification of the Skaergaard layered series.Geochimica et

Cosmochimica Acta, 38(10), 1549-1577.

Petry, C., Chakraborty, S., & Palme, H. (2004). Experimental determination of Ni

diffusion coefficients in olivine and their dependence on temperature,

composition, oxygen fugacity, and crystallographic orientation. Geochimica et

Cosmochimica Acta, 68(20), 4179-4188.

Pioli, L., Erlund, E., Johnson, E., Cashman, K., Wallace, P., Rosi, M., & Granados, H. D.

(2008). Explosive dynamics of violent Strombolian eruptions: the eruption of

Parícutin Volcano 1943–1952 (Mexico). Earth and Planetary Science

Letters, 271(1), 359-368.

Riker, J., (2005) The 1859 eruption of Mauna Loa Volcano, Hawaii: Controls on the

development of long lava channels. Unpublished M.S. Thesis, University of

Oregon.

Roberge, J., Wallace, P. J., White, R. V., & Coffin, M. F. (2005). Anomalous uplift and

subsidence of the Ontong Java Plateau inferred from CO2 contents of submarine

basaltic glasses. Geology, 33(6), 501-504.

Roedder, E. (1979). Origin and significance of magmatic inclusions. Bulletin of

Mineralogy, 102, 487-510.

Roggensack, K., Hervig, R. L., McKnight, S. B., & Williams, S. N. (1997). Explosive

basaltic volcanism from Cerro Negro volcano: influence of volatiles on eruptive

style. Science, 277(5332), 1639-1642.

Ruprecht, P., & Plank, T. (2013). Feeding andesitic eruptions with a high-speed

connection from the mantle. Nature, 500 (7460), 68-72.

Ruscitto, D. M., Wallace, P. J., Johnson, E. R., Kent, A. J. R., & Bindeman, I. N. (2010).

Volatile contents of mafic magmas from cinder cones in the Central Oregon High

Cascades: Implications for magma formation and mantle conditions in a hot

arc. Earth and Planetary Science Letters, 298(1), 153-161.

Ruscitto, D. M., Wallace, P. J., & Kent, A. J. R. (2011). Revisiting the compositions and

volatile contents of olivine-hosted melt inclusions from the Mount Shasta region:

implications for the formation of high-Mg andesites.Contributions to Mineralogy

and Petrology, 162(1), 109-132.

Page 185: FROM CINDER CONES TO SUBDUCTION ZONES: VOLATILE … · Ikuko Wada, Dominique Weis, Dan Rasmussen, Kathy Cashman, Katie Marks, and Philipp Ruprecht for their individual contributions,

169

Saleeby, J. B., Shaw, H. F., Niemeyer, S., Moores, E. M., & Edelman, S. H. (1989).

U/Pb, Sm/Nd and Rb/Sr geochronological and isotopic study of northern Sierra

Nevada ophiolitic assemblages, California. Contributions to Mineralogy and

Petrology, 102(2), 205-220.

Schiano, P. (2003). Primitive mantle magmas recorded as silicate melt inclusions in

igneous minerals. Earth-Science Reviews, 63(1), 121-144.

Sheppard, P. R., Ort, M. H., Anderson, K. C., Clynne, M. A., & May, E. M. (2009).

Multiple dendrochronological responses to the eruption of Cinder Cone, Lassen

volcanic National Park, California. Dendrochronologia, 27(3), 213-221.

Sherrod, D. R., & Smith, J. G. (1990). Quaternary extrusion rates of the Cascade Range,

northwestern United States and southern British Columbia. Journal of

Geophysical Research: Solid Earth (1978–2012), 95(B12), 19465-19474.

Straub, S. M., Gómez-Tuena, A., Zellmer, G. F., Espinasa-Perena, R., Stuart, F. M., Cai,

Y., & Mesko, G. T. (2013). The processes of melt differentiation in arc volcanic

rocks: Insights from OIB-type arc magmas in the central Mexican volcanic

belt. Journal of Petrology, 54(4), 665-701.

Suzuki, I. (1975). Thermal expansion of periclase and olivine, and their anharmonic

properties. Journal of Physics of the Earth, 23(2), 145-159.

Wallace, P. J., Kamenetsky, V. S., & Cervantes, P. (2015). Melt inclusion CO2 contents,

pressures of olivine crystallization, and the problem of shrinkage

bubbles. American Mineralogist, 100(4), 787-794.

Walowski, K. J., Wallace, P. J., Hauri, E. H., Wada, I., & Clynne, M. A. (2015). Slab

melting beneath the Cascade Arc driven by dehydration of altered oceanic

peridotite. Nature Geoscience, 8(5), 404-408.

Workman, R. K., & Hart, S. R. (2005). Major and trace element composition of the

depleted MORB mantle (DMM). Earth and Planetary Science Letters, 231(1), 53-

72.